Skip to main content

Full text of "The Electrolytic Dissociation Theory"

See other formats








THE ELECTROLYTIC 
DISSOCIATION THEORY 


BY 


Prof. R. ABEGG, Ph.D. 


4i in is idylls 


of the University of Breslau 




AUTIK 



CARL L. von ENDE, Ph.D. 

Assistant Professor of Chemistry , State University of Iowa 


FIR S T EDITION 
FIRST THOUSAND 


NEW YORK 

JOHN WILEY & SONS 
London : CHAPMAN & HALL, Limited 
1907 



Copyright, 1907, 

BY 

CARL L. von ENDE 



DEDICATED 


To My Dear Teacher and Friend 

Soante Qlrrljeuitw 

The Founder of the Dissociation Theory 




TRANSLATOR’S PREFACE. 


It seemed worth while to give English readers the 
benefit of this account of the electrolytic dissociation theory 
at the hand of a master of its details and applications. 

I gladly acknowledge my indebtedness, in so many 
ways, to my wife in preparing this translation, and also 
to Dr. Frederic Bonnet, Jr., of Worcester Polytechnic 
Institute, for helpful criticisms and suggestions. 

C. L. v. E. 


Iowa City, Iowa. 




AUTHOR’S PREFACE. 


As reasons for acting in place of the founder in pre- 
senting the dissociation theory, I may be permitted to 
mention the request of the publisher of this collection 1 
to undertake the task, and to this I would add the en- 
thusiasm which must seize upon every one who has taken 
the opportunity to study thoroughly this beautiful theory 
and learn how many old problems have been solved by 
it at one stroke, and how many new ones have come into 
view and been mastered. 

Quite recently, Roloff set himself the same task for a 
similar class of readers, and it offered some difficulty to 
keep this presentation from becoming merely competitive. 
I have for that reason touched but briefly upon the 
historical development, which is so adequately given by 
Roloff, and have endeavored to confine myself more to the 
detailed account of the chemical side, particularly to the 
development of the equilibrium relations among electro- 
lytes. While I believed that Ostwald’s exposition in his 
“ Foundations of Analytical Chemistry,” which showed 




“ Sammlung chemischer und chemisch-technischer Vortrage,” edited 
Prof, ©r, Felix. Ahrens (Ferd. Stuttgart). 

vii 



AUTHOR'S PREFACE 


viii 


so clearly the usefulness of the ion theory for every-day 
chemical purposes, was sufficient for the initiated, yet 
for a deeper insight into this attractive field a fuller account 
was desirable. In particular I have laid great stress on 
developing the formulae as simply and clearly as possible, 
and therefore the proofs have not infrequently been given 
in a form differing more or less from those in the original 
papers. These modifications have also seemed desirable 
as a result of my teaching experience. 

There is, of course, no intention of laying claim to new 
scientific achievements. We are justified, on the whole, 
in considering that Arrhenius himself has so thoroughly 
worked over the material that important advances of a 
general nature are scarcely to be expected; hence the 
nitiated will meet with new ideas and developments in 
only a few instances. 


Breslau, April, 1903. 



CONTENTS. 


PAGE 

Preface iii, v 

Fundamental Conceptions of the Theory x 

Mobility of the Ions 29 

Equilibria among Ions 37 

The Dissociation Constant 46 

Equilibria among Several Electrolytes 61 

Hydrolysis 76 

Avidity 95 

Indicators • 100 

Heterogeneous Equilibria 105 

Anomaly of Strong Electrolytes 121 

Influence of Pressure and Temperature on Dissociation. 134 

Non-aqueous Solutions 147 

Chemical Nature and Ionization Tendency of the 

Elements 159 

Index. 165 




THE THEORY OF 


ELECTROLYTIC DISSOCIATION. 


FUNDAMENTAL CONCEPTIONS OF THE THEORY . 1 

In the year 1887, when the Swedish physicist Svante 
Arrhenius propounded the theory of electrolytic disso- 
ciation (ionization), physical chemistry was passing from 
a kind of attractive side issue to a more central position 
of interest among chemists. Very interesting physical 
properties had been studied for some time and although 
the general laws were discovered, which with Ostwald 
we at present summarize under the caption Additive 
Properties, yet none of these furthered to any marked 
extent the constitution problems then prominent in the 
minds of organic chemists. The eyes of chemists were 
drawn again to the field of physical chemistry by the 
methods (discovered in 1883 by the recently deceased 


1 The sign * = positive ionic charge per equivalent, and the sign ' = 
negative ionic charge per equivalent. Chemical formulae in parenthesis, 
for example (H’)» indicate concentration of the kind of molecule in- 
closed, in this instance hydrogen ion.. 



2 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


Frenchman Raoult) for determining molecular weights 
by means of freezing- and boiling-points of solutions. 
This deter min ation of the molecular weights of substances 
in solution was an exceedingly useful aid in all kinds 
of chemical investigations. It was only natural, there- 
fore, that van’t Hoff’s 1 epoch-making theory of solutions, 
which appeared in the transactions of the Swedish 
Academy in 1885, and gave at once the theoretical ex- 
planation for the laws found by Raoult, should attract 
more attention in the chemical world than would have 
been possible under other circumstances, especially as it 
dealt with nothing less than judging the certainty of the 
conclusions based on the molecular-w r eight determina- 
tions of Raoult. General attention was further attracted 
to physical chemistry by the founding of the u Zeitschrift 
fur physikalische Chemie ” by Ostwald, at the beginning 
of his activities as a teacher at Leipzig, and by the com- 
pletion, shortly before that, of his well-known “ Lehrbuch 
der Allgemeinen Chemie,” in which he brought together 
and formulated as a whole all physical chemical knowl- 
edge. 

Through extended studies of his own on the conduc- 
tivity of electrolytic solutions, and through the theory 
of van’t Hoff as to the state of substances 2 dissolved in 
water, Arrhenius was led to look upon the so-called 
electrolytes, i.e., the acids, bases, and salts belonging 
especially to the field of inorganic chemistry, as broken 
up to a definite and usually large extent into their con- 

1 See “Sammlung chemischer und chemisch-technischer Vortrage,” 
Vol. V. 

Mbid. 



FUNDAMENTAL CONCEPTIONS OF THE THEORY . 3 


stituents, the ions. These ions, provided with electric 
charges, conduct the current by moving through the 
solution to the electrodes. 

This conception of free-existing parts of chemical 
molecules was nothing new among physicists. Davy 
(1808) and Faraday (1833), in their famous investigations 
on the electrolysis of fused and dissolved salts, and 
Hittorf, in his studies on the concentration changes caused 
by electrolytic conduction, had assumed and made prob- 
able such a molecular decomposition, even if such decom- 
position were not of very definite extent. 

These conceptions gained in significance through the 
experimental verifications of Buff (1855) and the theoret- 
ical proofs of Clausius (1857) and Helmholtz (1880), which 
showed that during electrolysis the components of the 
chemical molecules, though moving in opposite directions, 
do so without the least consumption of energy . The 
seemingly necessary assumption, on the part of chemists, 
of an affinity between the part-molecules was thus dis- 
proved, and there was no physical reason, therefore, for 
not considering the ions as independent of one another, 
that is, the molecules split up into such ions. We do not 
intend at this place to enter farther into the very inter- 
esting history of the subject, a detailed account of which, 
it may be well to mention, is to be found in the readable 
article of Roloff, 1 but rather to occupy ourselves with a 
presentation of the substance of the theory and its suc- 
cesses in the field of chemistry. 

As the name “electrolytic dissociation” indicates, the 


1 Zeitschr. f. angew. Chem., 15 , Heft 22-24 (1902). 



4 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

theory of Arrhenius includes all those substances which 
we term electrolytes, i.e., the substances which conduct 
the galvanic current in such a way that a movement of 
material masses takes place simultaneously in the direc- 
tions of the positive and negative currents. As has long 
been known, this peculiar kind of electric conduction is 
a property of salts, acids, and bases, and through them of 
almost all substances belonging to the field of inorganic 
chemistry. As Hittorf showed, we can formulate directly 
the statement that electric conduction is the essential 
characteristic of those substances known as “ salts” in 
the broader sense, and accordingly acids are looked 
upon as salts of hydrogen, and bases as salts of hydroxyl. 
Hereby a clear conception of “salt” was formulated for 
the first time, about which long experience had given us 
a practical but nevertheless inexact notion. 

In many cases the ions are determined by the nature 
of the products which separate at the electrodes during 
electrolysis; thus, for example, the ions of the saltCuCl 2 
are on the one hand the positive component Cu, separat- 
ing at the cathode, and on the other the negative Cl. 
Hittorf in his classical researches showed how one can in 
general determine the nature of the ions, that is, the com- 
ponents wandering in opposite directions, by the shift- 
ing of the concentrations which take place during elec- 
trolysis. That a salt, such as K2SO4, does not break 
up into K 2 0 and SO3, but into K 2 and SO4, is shown 
by comparison with KC1, in which Cl is the negative 
and therefore K the positive ion; and since both salts 
behave alike as to their positive component, having the 
K ion in common, the negative component of K 2 S0 4 



FUNDAMENTAL CONCEPTIONS OF THE THEORY . 5 


must be essentially SO 4, or the residue after taking away 
the K. 

On this conception as a basis one readily arrives at the 
long and vainly sought exact definitions of acids and 
bases. While it was sufficiently well known that their 
characteristic constituents were H and OH respectively, 
nevertheless it had not been possible to define under 
what circumstances these components showed acid or 
basic properties; for there are numerous compounds 
containing H or OH which are not necessarily acids or 
bases. The dissociation theory, however, defines these 
substances for us as such which contain H or OH in the 
'form of ions as the result of electrolytic dissociation, and 
makes clear at once the way of informing ourselves as to 
the degree of the acid or basic properties of a compound, 
by determining the concentration of these characteristic 
H" or OH' ions. This will be discussed later. 

. Let us here summarize a few of the more characteristic 
reactions peculiar to these two most important kinds of 
ions. 

The H* ions 

1. Change the color of u indicators”; for example, 

color blue litmus red, methyl orange red, 
decolorize red phenolphthalein solution -and 
yellow nitrophenol solution, etc.; 

2. Hasten catalytically the decomposition of esters 

by water into alcohol and acid, the inversion 
of cane-sugar, also the hydrolysis of maltose; 

3. Act as a solvent on many metals, marble, etc. ; 

4. Cause “acid” taste; 



6 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


5. Neutralize all characteristic properties of OH' 
ions. 

The OH' ions 

1. Change the color of indicators in the reverse 

sense of the H‘ ions; 

2. Act as saponifiers of esters; 

3. Accelerate catalytically the condensation of 

acetone to diacetone alcohol (also the reverse 
reaction), the conversion of hyoscyamine into 
atropine, and the disappearance of multi- 
rotation; 

4. Neutralize all characteristic properties of H # 

ions. 

Other ions have their specific reactions as well, but it 
is certain that, for reasons as yet unknown, we have to 
consider catalytic action as especially belonging to H’ 
ions and OH' ions, even if it is true that occasionally 
other substances can act catalytically. 

According to the theory of Arrhenius, these salts must, to 
a certain extent, be broken up into their ions, and the most 
convincing evidence for this conclusion was his discovery 
that all these salts were at the same time such substances 
as gave in aqueous solutions, according to the investi- 
gations of Raoult, freezing-point depressions which did 
not correspond to the molecular weights assumed from 
chemical considerations. While many substances dis- 
solved in water depress the freezing-point of the water 
by 1. 85° for each mole per liter, Raoult found that a 
considerable number of substances, particularly when 



FUNDAMENTAL CONCEPTIONS OF THE THEORY . 7 


dissolved in water, gave greater depressions ; or, according 
to the above-mentioned rule, they seemed to contain 
in a liter more than one mole, in spite of the fact that 
only one gram-molecule of the substance had been used 
for solution. Similarly the boiling-points of the same 
solutions showed too great a rise, thus urging the same 
conclusion. One was thereby brought to face the alterna- 
tive, either to doubt on the chemical assumptions the 
general tenability of Raoult’s law, or to admit on the 
basis of its validity that out of every molecule of these 
deviating substances several independent parts are 
formed. In assuming the latter the dissociation theory 
followed the same line of thought that was so successfully 
applied by Cannizzaro, Kopp, and Kekule in explaining 
the abnormal vapor densities of such substances as 
ammonium chloride, phosphorus pentachloride, and 
others. For, according to the van’t Hoff solution theory, 
the changes in the freezing- and boiling-points are the 
measures of the osmotic pressure of the dissolved sub- 
stances, and this osmotic pressure is entirely analogous 
to gas pressure. 

It is customary to speak of Raoult’s methods as methods 
for the determination of molecular weights; it would seem 
clearer, however, to call them methods for determin- 
ing molecule number or normal concentration, for the 
changes in freezing-point and boiling-point give directly 
only the number of moles of whatever kind contained in 
a definite volume of the solvent. Not until we consider 
the amount by weight contained in the solution do we 
arrive at the apparent molecular weight, which only 
represents a real molecular weight w T hen we can leave 



8 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

out of consideration the grouping together or splitting up 
of individual molecules. Since in so many cases this is 
hot* permissible, it is more rational to speak of the the- 
oretically unobjectionable molecular concentration given 
by the osmotic pressure or the methods of Raoult. The 
establishment of the qualitative agreement between the 
substances which conduct electrolytically and those which, 
according to the methods of determining normal con- 
centration, suffer a molecular splitting up was of very 
great immediate significance and was evidence supporting 
the idea of Arrhenius, for it was only natural to identify 
this molecular decomposition with the production of 
electrolytically conducting ions. The next consideration 
was the finding of a quantitative measure of proof. This 
was gained through Arrhenius, who considered that the 
degree of the conductivity must represent a measure of 
the ionic decomposition, in that the conductivity is 
essentially carried on by the ions and must take place 
the more readily the more ions are present, or the farther 
the electrically inactive molecules are split up into elec- 
tricity-transporting particles. Again, the molecular-num- 
ber methods (on condition that we look upon the ions as 
well as the undissociated molecules as independent indi- 
viduals) give a direct measure of the degree of ionic decom- 
position, so that the full molecular concentration of such a 
salt solution consists of that of its undissociated molecules 
increased by that of its ions, van’t Hoff had introduced 
a factor i into his theory of solutions, which indicates 
the number of times the molecular concentration given 
by the osmotic methods is greater than that to be expected 
from the chemical formula. 



FUNDAMENTAL CONCEPTIONS OF THE THEORY . 9 

Indicating by a the fraction of a mole 1 of a salt which 
is split up into ions, and by (i — a) the undecomposed 
portion, we can calculate the factor i if we know n the 
number of ions into which one molecule can break up. 
We have then, for i mole, the part (i — a) left undis- 
sociated and (n-a) ionic molecules formed from the 
rest; the sum total of undissociated and ionic mole- 
cules is therefore (i — OL-\-na) individuals, so that 
i=i — a-\-na = i + (n — i)a. 

A first proof of the theory is given by the fact that a , 
the degree of dissociation, can be derived from the measure- 
ment of the conductivity. Under the assumption, which, 
as we shall find later, holds for neutral salts, that at very 
great dilutions the breaking up of the salt into ions 
becomes practically complete, the comparison of the 
conductivities produced by one mole of the salt when 
dissolved in a definite volume of water with the con- 
ductivity it assumes at very great dilution gives the degree 
of ionic decomposition. While we shall later consider 
in detail the more exact determination of the degree of 
dissociation of different substances, let us here anticipate 
to the extent of saying that Arrhenius, in the year 1888, 2 
in testing on an extended scale the relationship between 
i and a , as derived above, found an excellent substan- 
tiation of the theory. 

A very important question, the solution of which had 
occupied chemists in vain for a long time, was this, What 
is formed in a mixture of salts? For instance, to what 

1 Mole = gram-molecule. 

2 Zeitschr. physik. Chem., 2 , 491 (188S). 



10 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

extent, if at all, does a reaction take place to form K0CO3 
and Na2S04 when one mole of K2SO4 and one mole of 
Na2C03 are brought together in solution? It is sur- 
prising to note that it is by no means generally known 
that the solution of the two salts named is identical with 
the one obtained by mixing a mole each of K2CO3 and 
Na 2 SC>4. The author has repeatedly met chemists who 
to this day in all seriousness discuss how the metal and 
acid constituents of such a mixture are mutually combined. 

The dissociation theory, however, gives for this an 
entirely convincing explanation that can readily be tested 
at the hand of experience. Since, according to this 
theory, K 2 S 0 4 and Na 2 C 0 3 as well as K 2 C 0 3 and 
Na 2 S 0 4 are to a large extent split up into the ions K, Na, 
S 0 4 , and C 0 3 , it is clear that it can make no difference 
from what solid substances these ions take their origin, 
for in the solution they have become independent of the 
constituent originally combined with them. This con- 
sideration is of practical significance, for example, in the 
artificial preparation of mineral-water salts which shall 
give solutions identical with those of the natural springs . 1 
Suppose analysis shows that a certain well-water contains 
for 1 equivalent of sulphate 2 equivalents of chlorine, | 
equivalent of potassium, and 2 \ equivalents of sodium, it 
is absolutely immaterial and leads to exactly the same 
solution if we mix \ equivalent K 2 S 0 4 , \ equivalent 
Na 2 S 0 4 , and 2 equivalents NaCl, or 1 equivalent Na 2 S 0 4 , 
1 J equivalents NaCl, and \ equivalent KC 1 , or in general 
any quantities of the four salts made up of the four 


1 Zeitschr. f. Elektrochem., 9 , 185 (1903). 



FUNDAMENTAL CONCEPTIONS OF THE THEORY, n 

components, provided we meet the condition that the total 
amount of K, Na, SO 4, and Cl equals that of the analysis. 
This experimental fact may be summed up in the state- 
ment that salts are such substances as are in a high 
degree subject to a so-called mutual decomposition, 
which, and this is of importance, takes place with im- 
measurable velocity. 

An exceedingly important fact of chemistry and one in 
very close relationship with the above is the striking 
phenomenon that in practically all salts the basic and 
acid components show exactly the same reactions no 
matter in what combination these components happen to 
be. Thus it is a well-known fact that all soluble barium 
salts give with all soluble sulphates one and the same 
reaction, that is, form barium sulphate. Similarly 
copper is precipitated as copper sulphide by hydrogen 
sulphide from all of its salt solutions quite independent of 
the acid component with which it is combined; chlorine 
gives the same precipitate with silver nitrate no matter 
whether it is contained in KC 1 , NaCl, CuCl2, etc., etc. 
On the other hand, in the case of organic compounds, the 
same radical at times shows greatly varying reactions, 
depending on the nature of the other elements combined 
with it. Now it seems highly improbable and directly 
contradictory to the character of chemical compounds 
that different compounds should give an identical reaction 
with, the same substance; yet in the case of salts, as we 
saw above, we cannot avoid this very conclusion. But 
here again the dissociation theory offers the solution of 
the dilemma, for, according to its concept, the same radical 
in tlie different salts, in consequence of . electrolytic dis- 



12 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


sociation, appears as a free and, in all cases, equal ion 
and therefore gives the same reaction. The entire 
structure of analytical chemistry is built up on this fact, 
and the especial feature of the system of inorganic analysis 
is its relative convenience and simplicity, which is condi- 
tioned essentially on the identity of a substance being 
maintained in spite of its manifold combinations. On 
the contrary, an organic system of analysis must be, to all 
intents and purposes, counted with the impossibilities on 
account of the infinite diversity of the reactions. 

A further peculiarity of a salt solution is the additive 
nature of its physical properties, such as color, density, 
refractivity, conductivity, and so on. By this we under- 
stand that these properties can be made up of two quan- 
tities, one of which can be assigned to the base alone, the 
other to the acid alone, so that if we know these separate 
values for a certain number of radicals, we can calculate 
the properties of each combination by simply adding the 
corresponding quantities. As a type of the additive 
properties of electrolytes we can take that of the chemical 
reactions just discussed. 

Thus Ostwald 1 established the fact that the character- 
istic absorption spectra of equivalent solutions of per- 
manganates, for instance, are the same, independent of 
the (colorless) cathion with which the Mn0 4 was 
combined, that is to say, each ion imparts to the solution 
its own peculiar color. For this reason all dilute solutions 
containing copper ions are blue, all ferrous salt solutions 
greenish, all rosaniline salts red, etc. We may further 

g - ' — “ — — 

. r 

1 physik, Cheni./ 9 , $79 (1893) 


FUNDAMENTAL CONCEPTIONS OF THE THEORY. 13 

conclude that all ions present in a colorless solution have 
no color of their own: as H*, K*, Na*, Li*, Ba**, Sr", 
Ca**, Mg**, Be**, OH', F', Cl', Br', I', SO/', N 0 3 ', 
CIO/. 

Frequently the color can also give interesting informa- 
tion as to the constitution of inorganic salts. Thus neither 
potassium ferrocyanide nor potassium ferro-oxalate pos- 
sesses the green color of the Fe*‘ ion, but they are yellow 7 
and red respectively, and hence must contain the Fe 
in some other form, that is, as the complex ions Fe(CN)e / ' // 
and Fe(C 2 0 4 )2" respectively, as was demonstrated by 
Hittorf. Likewise the change in color of Cu" ions by 
ammonia, or their decolorization by potassium cyanide, 
discloses the fact that complex ions are formed in which 
the copper is no longer present as Cu*\ All such con- 
clusions have found their remarkable confirmation 
experimentally. 

Buckingham found 1 that fluorescence is often es- 
sentially the property of an ion and is wanting in the 
undissociated substance, as in the case of eosine, / 9 -naph- 
thylamine disulphonic acid (1:2 :5), and quinine. Here 
the ions retain their entirely independent properties. 

Further, Valson found that equivalent solutions of 
KC 1 and NaCl show a difference in specific gravity, 
which remains unchanged when w r e substitute for Cl 
any other acid residue, thus indicating that the difference 
is independent of the nature of the acid. In the same 
way any two acids give a constant difference independent 
of the basic constituent. At every hand, then, we have 


\ geitschTr physik,- Chein., 14 , 129 (1894), 



14 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

evidence that the components of electrolytes do not 
mutually influence each other. 

All these phenomena are to be looked upon as necessary 
consequences of the ionic dissociation, .for the properties 
of the ions must be constant as long as the ions remain 
the same. If therefore the combination of the ions does 
not affect their nature, that is, leaves them independent 
and free, the additive nature of the properties follows as 
a necessity. 

A great number of objections have been raised to the 
conception that the constituents of “ salts ” in the broader 
sense lead a chemical existence independent of one 
another. Above all, the opinion had always been held 
that the foremost salt-formers, the alkalis on the one hand 
and the halogens on the other, were bound one to the 
other by extraordinary affinity forces, since they react 
with very great affinity manifestations, such as intense 
heat liberation and even light. And now would these 
components be separated again by simple solution in 
water? In asking this, the fact was entirely overlooked 
that the dissociation theory does not assume that the 
electrolytes split up into the atoms or molecules from which 
they were formed, but that these decomposition products 
are essentially different from those atoms or molecules, 
in that they are electrically charged. These charges are 
of enormous magnitude, since, according to Faraday’s 
law, each ion carries for its formula weight, in grams, 
96580 coulombs per equivalent. 

It is also claimed that the abnormal osmotic pressures 
of the electrolytes can be explained by a hydrolytic de- 
composition— for instance, . NaCI +H 2 0 =NaOH +HCL 



FUNDAMENTAL CONCEPTIONS OF THE THEORY. 15 


This assumption, however, leads ad absurdum, since, in 
the first place, the assumed decomposition products, 
NaOH and HC 1 , which in their turn cannot undergo 
further hydrolysis, also show, like the salt, too high an 
osmotic pressure. And secondly, as has long been 
known, solutions of acids and bases, which according 
to the assumption of opponents would have to exist 
alongside without reacting, on the contrary, do react 
with one another very energetically. The theory of 
neutralization here involved will be further discussed 
later on in the light of electrolytic dissociation. 

Another criticism has been the impossibility of applying 
to ionic decomposition the crucial experimental test of sep- 
arating the decomposition products of the split-up body, as 
in the case of the gaseous dissociation of ammonium chlo- 
ride. In this it was overlooked that the separating of the 
oppositely charged ions cannot take place to a measurable 
extent by reason of these very charges, since it would 
require the setting free of enormous quantities of elec- 
tricity. For should w T e wish to isolate from each other 
only one milligram equivalent of cathion and anion, it 
would be necessary to have appear, at different points of 
the system in space in which this separation was to take 
place, electrostatic charges of 96 coulombs. This means 
charges of a magnitude sufficient to give to a large flask 
provided with a condenser covering, such as was employed 
by Ostwald and Nemst, 1 a potential of about 8000 volts! 
Nernst, 2 however, showed, in his epoch-making theory 


1 Zeitschr. physik. Chem., 3 , 120 (1S89), 

’Ibid,, 2 , 01, 3 (:§§§); 4 , i?9 (1889), 



l6 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

of the diffusion of electrolytes and the so-called diffu- 
sion chains, how these local separations of cathions and 
anions, even though immeasurably small, do take place, 
and can be employed for a quantitative calculation of 
these changes. Again, together with Ostwald 1 he 
demonstrated, by using two vessels connected with a 
siphon filled with an electrolytic acid solution, that it is 
possible by means of most powerful electrostatic in- 
fluences to transfer a sufficient excess of ions with 
positive charges into the one capillary vessel, and nega- 
tively charged ions into the other, so that upon conduct- 
ing away the excess of electricity in the former the dis- 
charged hydrogen ions are made visible as bubbles of 
hydrogen gas. Hence this objection can also be 
considered as being in every particular effectively refuted. 

Likewise another theory, known as the hydrate theory, 
attempted to meet the phenomena explained by the theory 
of Arrhenius, and in particular to explain the important 
phenomenon of the increased osmotic pressure. This 
theory states that the molecules of the dissolved sub- 
stances combine with considerable quantities of the 
water solvent to form hydrates, whereby the molecular 
concentration of the dissolved substance, i.e., the ratio 
of the number of dissolved molecules to that of the 
free uncombined solvent molecules, may appear greatly 
increased in that the molecules of solvent consumed for 
hydration no longer act as solvent. There is nothing to 
be said against the fundamental conception of this theory 
of a chemical union between the two components of a 


t g^tsch'r. phvsik, Ckem., 3 , 196 (188$).' 




FUNDAMENTAL CONCEPTIONS OF THE THEORY . 17 


solution; on the contrary, the results of recent and varied 
physico-chemical research make it appear more and 
more probable. In spite of this the hydrate theory is 
incapable of competing from a quantitative standpoint 
with the dissociation theory. Since the abnormally high 
osmotic pressures also appear in extremely dilute solu- 
tions, in fact are most evident there, the hydrate theory 
would have to assume, in case of a tuW normal solution 
which gives double the normal freezing-point depression, 
that, of the approximately 55 moles of water contained 
in one liter, about one half, or 27 moles of water at least, 
are bound to yioVir mole of the dissolved substance, 
giving as a formula for this hydrate 1 mole salt + 2 7000 
H 2 0 . Further, it is evident that, in spite of the varying 
concentration of the solute, the number of bound water 
molecules would always have to remain approximately 
constant, provided the abnormality factor i of the osmotic 
pressure, as is often actually the case, scarcely varies with 
the concentration. This conclusion is altogether contrary 
to the law of mass action, according to which the hydrated 
portion of the salt must be proportional to the product 
of the anhydrous portion and the active mass of the 
water, expressed by the equation: 

Hydrate = k • (Anhydride) • (Water). 

Now on account of thermodynamical reasons (Nemst) 
the active mass of water is proportional to its vapor 
pressure, and this, according to Raoult’s measurements, is 
only about 2% smaller for a normal solution than for 
pure water, i.e., for dilute solutions it may be considered 
practically identical with that of water, so that the active 



1 8 the theory of electrolytic dissociation. 

mass of the water in the above equation is -constant; 
which means that the quantity of hydrate in such solu- 
tions is proportional to the quantity of anhydride. It 
follows that the quantity of water bound as hydrate would 
have to become less and less with increased dilution of 
the solute, and so the abnormalities of the osmotic pressure 
noted at greater concentrations would also continually 
decrease, which is directly contrary to the observed facts. 
But leaving all of this out of the question, the dissociation 
theory is capable of giving in an extremely convincing 
manner orientation as to the magnitude of the abnormality 
factor i according to the number of ions into which an 
electrolytic molecule splits up, in that binary salts of the 
type of KC 1 can give rise to twice the normal value of the 
osmotic pressure, ternary salts such as K 2 S 0 4 or MgCl 2 
to three times, and so on. Thus, for instance, we can 
read directly from the formula K 4 Fe(CN) c that in 
consequence of the decomposition into five ions, 4K and 
the anion of the tetrabasic hydroferrocyanic acid, the 
maximum molecular osmotic pressure (at greatest dilution) 
must be five times the normal; for sodium mcllitate, 
which can split up into seven ions, Taylor 1 attained nearly 
the maximum value (see table, p. 25). On the other 
hand, from the value of the factor i, the hydrate theory 
would have to set up an hypothesis as to the degree of 
hydration for each particular salt concerned, but this 
hypothesis would be encumbered by the previously 
mentioned defect. So we can hardly be in doubt as to 
which theory to prefer, especially when we consider that 


1 Qstwald’s Zeitschr ., 27 , 361 ( 1898 ), 



FUNDAMENTAL CONCEPTIONS OF THE THEORY . 19 

the hydrate formulae, which it would be necessary to 
employ in order to explain the osmotic pressures, do not 
in the remotest agree with the known water of crystalliza- 
tion formulae, making them seem altogether arbitrary. 
Even if we can herewith consider this theory as disposed 
of, so far as explaining the fundamental facts of dissocia- 
tion is concerned, we shall nevertheless meet the same 
again later on (p.127), where for certain anomalies of 
electrolytes it offers a possible explanation. 

The most prominent problem which the dissociation 
theory had to solve — its fundamental concept once accepted 
— was the determination of the degree of dissociation of 
the different electrolytes. It has been mentioned that 
this may be done by means of the abnormality of the 
osmotic pressure, by introducing into the calculation the 
increase in the number of molecules produced by the ions 
formed in extremely dilute solutions, where we may 
consider the ionization as complete and the abnormality 
factor i must reach its maximum limiting value. This 
value at the same time indicates the number of ions that 
are formed from the salt molecule. 

Another way to get at the degree of dissociation 
Arrhenius found in the study of electric conductivity. 
The specific conductivity k of an electrolyte is the current 
strength which flows when the same is placed between 
two electrodes of 1 square centimeter area, 1 centimeter 
apart, with a potential difference of 1 volt. 

For one and the same electrolyte this specific conduc- 
tivity is naturally very much dependent on the concen- 
tration, since as it varies, the amount of the electrolyte, con- 
tained in the 1 centimeter cubed between the electrodes. 



20 the THEORY OF ELECTROLYTIC dissociation. 

must vary. The study of the specific conductivity can 
therefore give directly no means for finding out in what 
way the molecule of the substance changes its capacity 
for conducting electricity with varying dilution. Such 
a means is gained, however, from the specific conductivity, 
if we reduce the same by calculation to one and the same 
concentration — for instance, to one equivalent in a cubic 
centimeter; or if we imagine (Ostwald) the use of elec- 
trodes, which remain at a fixed distance apart of i centime- 
ter, but which with increasing dilution of the electrolyte 
always increase in area, so that the volume of liquid 
included between the electrodes always contains just one 
equivalent of the electrolyte. The conductivity of one 
equivalent in its varying dilutions, thus observed, is 
evidently a magnitude capable of giving information as to 
the change of the molecular condition, in so far as this 
influences the conductivity. Indicating the equivalent 
conductivity by A, and the concentration, in equivalents 

K 

per c.c., by tj} then A = — . 

Now Kohlrausch had found that the equivalent con- 
ductivity A increased with increased dilution for all 
electrolytes and in many cases approached a limit value 
A 0 for very great dilution. This best attainable con- 
ductivity A 0 the theory of Arrhenius conceives as belonging 
to that molecular condition which consists essentially 
of ions, so that it can devote itself entirely to the trans- 
portation of current, while at higher concentrations the 

1 Concentration in normals c = equivalent /liter stands to this in the 

. c 

ratio — =1000. 

n 



FUNDAMENTAL CONCEPTIONS OF THE THEORY. 21 


values for A (< A 0 ) are characteristic of the extent to 
which the molecule is ionized. 

If this conception is correct, there must exist a very 
simple law for the A 0 values of different electrolytes when 
we consider the manner in which these are dependent on 
the nature of the ions. Suppose we indicate by u the 
velocity given to any cathion by the potential fall of i 
volt per centimeter, by v the corresponding value for an 
anion, and recall the fact that, according to Faraday’s 
law, the charge of F coulombs carried by each equivalent 
of any ion is always equal, then in one second there will 
be moved between the electrodes mentioned above u-F 
coulombs by the cathions in the positive direction, and’ 
simultaneously v-(—F) couloinbs by the anions in the* 
negative direction, that is, the total current flowing wall, 
be 

A 0 =u-F— v-(— F)=u-F+v-F=(u±v) -F COU ^- 

sec. 

or amperes. Since u and v depend entirely on the nature 
of the ions, it follows that the A 0 values of different 
electrolytes must be purely additive, i.e., composed of 
factors characteristic of the two ions, so that, for instance, 
the differences for K and Na salts should be exactly the 
same whether derived from the chlorides, nitrates, etc., 
for 

(U}H-\-Va)F— (Uiz a -h‘VcdF=(Uj£ + Viio)F — (^Na + ^NO^F 

= (U'K. — U Na)F. 

This relationship was in fact discovered by F. Kohl- 
rausch in 1876, and is called the law of the independent 



22 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


migration of the ions. This law, which, as the formula 
shows, does not directly give individual specific ionic 
velocities but only the differences of two such, may be 
illustrated by the following small table, which includes the 
figures of those K and Na salts whose acid radicals are 
given in the first column : 


K ' "■ 

J =— for iooo n — c=o.gq£>i Eqtjiv. /Liter. 
V 

(Kohlrausch, 1900 and 1885 ) 



K 

Na 

(u K ~u Na )F 

Cl 

129 .1 

108 . 1 

21.0 

no 3 

12 5-5 

104 . 6 

2O.9 

io 3 .. 

97.6 

76.7 

20.9 

(so 4 )^ 

133-5 

no. 5 

23.O 

£(*ci -’to,)- 

F *so, -*ro 3 i 

F ^o 3 - v io 3 ) 

3 1 -5 

35-9 

27.6 

3 1 -4 

33-8 

27.9 



For a large class of electrolytes, namely, almost all 
salts as well as the strong acids and bases, the values of 
A 0 may be obtained by direct measurement, since with 
increasing dilution the values of A show clearly a con- 
vergence toward a limiting value, as can be seen from 
the following series for KC 1 (i8°) (Kohlrausch, 1885). 


Potassium Chloride (i8°). 


C= I 

0. 1 

0.01 

o.oor 

0.0001 

0 

A — 98.2 

in. 9 

122.5 

127.6 

129.5 

131.2 


AA = 13.7 10.6 5.1 1.9 

A similar series for acetic acid 

A = 1.32 4.6 14.3 41 107 

3.3 9.7 26.7 66 



FUNDAMENTAL CONCEPTIONS OF THE THEORY. 23 


gives no evidence of such a convergence in dilutions 
experimentally accessible, as is typical of all weak electro- 
lytes. In such cases, however, A 0 can be obtained in- 
directly by means of Kohlrausch’s law, by making use of, 
for instance in the case of acetic acid (H acet.), the ex- 
perimentally accessible A 0 values for K acet., KC1, and 
HC1, and calculating as follows: 


^0 (K acet.) + A 0 (H C1) ~ ^0 (KC1) = , 


for 


* 

F[(Uk+ V ace t.) + (u s + v a) — ( M K+^a)] = -f’(^H s +^acet.); 
in short, 

K+acet. +H+C1— K— Cl=H+acet. 


Or in words, we begin with A 0 of a salt of the weak acid 
and add to it the difference between the A 0 values of a 
strong acid and its salt, which has the same cathion as the 
salt of the weak acid. 

A prime criterion of the correctness of the course of 
reasoning lies in the conclusion that the equivalent 
conductivities A 0 for “ infinite dilution,” calculated with 
the aid of Kohlrausch’s law, must under all circumstances 
be greater than the experimentally determined equivalent 
conductivities A of weak electrolytes; for if A differs 
from A 0 , this can only be in the direction corresponding 
to an incomplete ionization, because A 0 is necessarily 
associated with complete ionization. 

For concentrations in which all the molecules do not 



24 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

split up into ions, the degree of ionization being less than 
i, the equivalent conductivity will also have to be smaller 
than A 0j for if there are present per equivalent only a ions, 
a indicating the degree of ionization, then for each equiv- 
alent only (analogous to our findings for A 0 , p. 21 ), 

, , . „ coul. 

A =a(u+v)F ^ 

can be carried. Substituting from the above equation 
the value for A 0y 


A 



the sought-for new definition of the degree of ionization 
(dissociation), determinable by electrical means. The 
above consideration, that the measured A values must 
always be less than the A 0 values calculated by means of 
Kohlrausch’s law, is confirmed without exception by ex- 
perience. We are therefore justified in building further 
on this foundation and in looking upon the ratio of the 
equivalent conductivity A , of a particular concentration, 
to A 0 (exterpolated or calculated as above) at infinite 
dilution, as the direct measure of the degree of decom- 
position into ions, and in formulating, as did Arrhenius, 
the equation for the degree of ionization 



The dissociation theory withstood the first crucial test 
in that a , the degree of ionization calculated from the 



FUNDAMENTAL CONCEPTIONS OF THE THEORY. 25 


equivalent conductivities, showed such surprising agree- 
ment with that given by the deviations of the osmotic 
-'pressure according to the formula (p. 9) 

i=n-a + (1 — a) = i~r(n— 1) °l 
or 


i —1 


a 

71 — 1 


(2) 


The following figures, taken from freezing-point deter- 
minations of Arrhenius and others, show this agreement: 


Comparison of the Osmotically and Electrically Measured 
Abnormality Factors. 

(van’t Hoff and Reicher, 1889.) 


Salt. 

Concentra- 

tion. 

♦ 

(osmot.). 

i 

(freez.). 

i 

(electr.). 

KCI 

0. 14 

1 .81 

_ 

1.86 

NH 4 C 1 

00 

"tf- 

H 

o' 

I.82 

— 

I.89 

Ca(N 0 3 ) 2 

0.18 

2.48 

2-47 

2.46 

K 4 Fe(CN) 6 

°- 3 S<> 

3-°9 

— ■ 

3-°7 

MgSO, 

O.38 

1-25 

1.20 

i -35 

LiCl 

O.I3 

1.92 

i -94 

1.84 

SrClj 

O. l8 

2.69 

2.52 

2-51 

MgCl*. 

O. 19 

2.79 

2.68 

2.48 

CaCl 2 

O. 184 

2.78 

2.67 

2 .42 

CuCLj — 

0.188 

— 

2.56 ; 

2.41 

Na 6 C 12 0 12 

O.OOl8 


5 * 9 2 

~ 


With this there were at hand two methods, differing 
in principle yet giving like results for getting at the dis- 
sociation relations of the long list of electrolytes. The 
results of these investigations, carried out by Arrhenius 
and the Leipzig School under the leadership of Ostwald. 
may be summarized as follows: 




26 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

1. Strong electrolytes are such salts, acids, and bases 
which even in considerable concentrations ionize very 
much more than half and contain as cathions any alkali 
metal (Cs, Rb, K, Na, Li), or as anions one of the acid 
residues N0 3 , CIO3, CIO4; furthermore, combinations 
of the following cathions and anions in so far as they are 
soluble: 

r NCV 

NH 4 -, Ba”, Sr", Ca” Mg", 1 F, Cl', Br', I' 

Mn", Zn", Fe*‘, Co", Ni”, with S0 4 ", S 2 0 6 " 
Pb-, H*, Hg 2 **, Ag* Cr0 4 ", Cr 2 0 7 " 

CClsCOO' 

Accordingly we have belonging here all alkali salts, 
nitrates, chlorates, perchlorates, as well as the strong 
acids HC1, HBr, HI, HNO s , H 2 S0 4 , H 2 F 2 (the last two, 
it is true, are markedly less ionized than the previous ones) 
also all sulphonic acids, and practically speaking all solu- 
ble neutral salts, the ammonium and substituted ammonium 
salts inclusive; of the bases, that is, the hydroxyl com- 
pounds, we have only the tetra-substituted amine bases, 
while ammonia and the substituted amines up to. the 
tri-substituted belong to the next class. 

2 . Weak electrolytes include first of all the three large 
classes of the organic carboxylic acids, phenols, and 
primary to tertiary substituted amine bases; also am- 
monia, and the following compounds which form an 
exception to the other neutral salts: 

CdCl 2 , CdBr 2 , Cdl 2 „ HgCl 2 , Hg(CN) 2 , Fe(CNS) 3 , 

FeF 3 , Fe(acet:) 3 , 



FUNDAMENTAL CONCEPTIONS OF THE THEORY . 27 

and the weak inorganic acids: 

H 2 S, HCN, H3BO3, H3PO0, H3PO3, H3PO4, 
H 2 C 0 3 , H 2 SO s , HoSeOs, HN 0 2 , HCIO, HIO3, HI 0 4 - 


A large number of inorganic salts not included in the 
above, such as those of the last-named acids, show the 
phenomenon of hydrolysis and will receive special men- 
tion later (p. 92). 

Between these two extreme classes of electrolytes we 
have, of course, all transitions, for the two classes are only 
gradually differentiated, since, as has been mentioned and 
as will later be discussed in detail, the degree of ioniza- 
tion is greatly dependent upon the concentration. A 
kind of transition class, designated as electrolytes of 
medium strength, might be set up, consisting on the one 
hand of the salts of the heavy metals, and on the other of 
the strongest carboxylic acids, such as tartaric, citric, 
oxalic, and formic, also many halogen and nitro- sub- 
stituted carboxylic acids. 

For reasons to be mentioned later, a very special interest 
attaches to the extremely weak electrolytes, which are 
transitions from electrolytes to chemical compounds 
incapable of electrolytic dissociation. These will be dis- 
cussed in a chapter to follow, and figures given which are 
characteristic of them; here it will be sufficient to say 
that they include hydrocyanic acid, hydrogen sulphide, 
boric acid, carbonic acid, phenol, and above all waters 
also the bases aniline, pyridine, etc. (see table, p. 53). 

In addition to the conductivity method for determining 
the degree of ionization we have another electrical method, 



28 the theory of electrolytic dissociation. 

based on the Nemst theory of concentration chains. 
x\ccording to this the electromotive force of such chains, 
in which the same electrode metal dips into a concentrated 
and a dilute solution, is proportional to the logarithm of 
the concentration ratio of the metal ions in the two 
solutions. Since the mathematical form of this function 
requires extremely accurate measurements in order to 
determine small differences of ionic concentrations, this 
method has not been applied for this purpose until recently 
by Jahn (see p. 120), though it has been employed with 
great success in recognizing the extremely small ionic 
concentrations of very difficultly soluble electrolytes. 



MOBILITY OF THE IONS. 


The figures enumerated on p. 22 show that the con- 
ductivity differences between alkali salts of the same 
acid are independent of the nature of the acid, that 
is, they are evidently dependent only on the difference of 
the cathion. In a similar way we get equal differences 
for a change of the acid constituents, no matter from what 
alkali salts we take the conductivities. It is necessary 
to take only one step further in order to determine, in the 
case of any salt, how the conductivity is divided between 
the anion and cathion, and to calculate from the above 
table the part each ion takes in the conductivity. This 
was done in the classical researches of Hittorf on the 
concentration changes in the vicinity of the electrode 
during the electrolysis of salts, and permits drawing a 
conclusion as to the apportionment for the two ions in 
transporting the current. 

Suppose we conduct a definite amount of electricity, 
say 96580 coulombs, the quantity carried by one ion 
equivalent, through an electrolytic cell, then 

1. At each of the electrodes, according to Faraday’s 
law, one equivalent of the respective ions is separated. 
This we shall look upon either as remaining in solution, 
as is actually the case in the electrolysis of salts such 
as KgSO^ or ? in case it is precipitated, as belonging 



30 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


to the solution immediately surrounding the particular 
electrode. 

2. It is necessary for us to gain some insight into the 
mechanism of the current transport in so far as it takes 
place within the solution, i.e., between the electrodes. 
Of our 96580 coulombs one part is carried by the positive 
ions and the other part by the negative ions in their 
wandering to the electrodes, so that a certain quantity of 
anions move away from the cathode and a certain quantity 
of cathions away from the anode and will be wanting at 
their former places. It can readily be seen that this 
reduction of the concentration at the electrodes must give 
a measure of the nature of the ratio of the rates of migra- 
tion of the ions wandering in opposite directions. For 
example, in the case of equal mobility, that is, like rates 
of wandering for cathions and anions, the reduction in 
concentration of these ions at the electrodes from which 
they migrate must be exactly equal; with unequal mobility 
the reduction in concentration must be greater at the 
electrode from which the ion of greater velocity moves. 
The ratio of these ionic concentration reductions at both 
electrodes, as measured by Hittorf, represents, in other 
u 

words, the ratio in which the transport of current dis- 
tributes itself between both ions. Calculating from the 
shifting of the concentration of one of the ions the cou- 
lombs carried by the ionic matter transported away, and 
comparing this with the total coulombs (measured, for 
example, by means of a voltameter) which flowed through 
the electrolyte during the time in which the measured 
concentration change took place, we obtain the so-called 



MOBILITY OF THE IONS. 


3 1 


transference number, which represents the fraction 
u j v \ 

— ; — ( or — ; — I of the total number of coulombs trans- 
u+v \ u+vj 

ported by this ion. The transference number i for 
the cathion of a given salt would mean that the entire 
transportation of current was carried on by the cathion, 
while the anion had no part in it whatsoever. This, 
however, could take place only in the extreme and non- 
existent case of the mobility of the anion being infinitely 
smaller than that of the cathion, for, as both ions are 
moved by the same electrical driving force, their velocity 
must be proportional to their mobility. The transference 
number 0.5, which, on the other hand, is not infrequently 
found at least approximately, would indicate that the 
electric current is carried half by the cathion and half 
by the anion, or that both ions possess the same mobility. 
The most exact measurements of transference numbers 
made are those with potassium chloride, which give for 
the ion K the value 0.497. 

Now we know' from Kohlrausch’s measurements that 
for KC 1 (see p. 22) 

A 0 = (u -f ^96580 = 1 30.1 , 
and with the aid of the second equation, 


u 

u+v 


0.497, 


we are in position to calculate separately the values for 
u and v. In order to avoid too small values, it is better 
to employ, instead of the “ absolute mobilities ” u and v, 



32 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

the “ electrolytic mobilities ” Ik and l A for cathion and 
anion, which are 96580 times greater. We may then 
write 

A 0 =l K +l A = 96580(11 +v)=i 30.1 

and 

U 

U + V Ik+Ia 

from which we get for the potassium ion 
l K = 64.67, 

and for the chlorine ion 


/ A = 65.44. 


These values at once enable us to get at the mobility 
of other ions by using the equivalent conductivities of 
other potassium salts and other chlorides, and subtracting 
from these the mobility values of K* and Cl' respectively. 
What is more, we are also in position to calculate the 
maximum equivalent conductivity for such electrolytes 
whose measurements do not show any such maximum 
conductivity. 

For example, in order to obtain the electrolytic mobility 
of any anion, say F', it is only necessary- to know A 0 for 
KF, which Kohlrausch (1902) found to be 




Subtracting from this Ik = 64.67 
we have left / FW ne= 46.68 



MOBILITY OF THE IONS. 


33 


The mean values of the most accurately known elec- 
trolytic mobilities 1 at i8° are given (according to Kohl- 
rausch 2 3 ) along with their temperature coefficients a in 
the following tables: 



1 The l values give the conductivity of i mole of the ion in i c.c. (not 
in i /!). 

2 Berl. Akad. Ber., 26 , 586 (1902). 

3 For the H* and OH' ions the electric mobilities are known with a 
much less degree of certainty, since it is impossible to follow up the A 
values of acids and bases to the very great dilutions where the con- 
ductivity of the water and its unavoidable impurities play a part not 
yet determined. 






34 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


These numbers enable us to calculate by addition A 0 , 
the equivalent conductivity at infinite dilution, for all 
salts formed by combinations of the above ions. They 
have been amply confirmed by the fact that the trans- 
l 

ference numbers ; — — calculated from these values 

agree admirably with those found. 

Evidently an exceptional position is held by the ions 
of water, H* and OH', of which the former is about five 
times and the latter about three times as mobile as the 
most mobile of its kind. In consequence of this, among 
the strong electrolytes the acids and bases (comparing 
equivalent solutions) are much better conductors than 
all neutral salts. 

For electrolytes with ions of greater valence the relations 
are more complex, in that the values for A do not converge 
sufficiently at convenient dilutions to give accurate values 
for Aq. However, as has recently been shown by Steele 
and Denison , 1 in the case of such electrolytes the trans- 
ference numbers, which vary considerably with the con- 
centration, converge toward values showing a good 
agreement with the A 0 values measured by Kohlrausch. 

A regularity in the magnitude of the mobilities may be 
formulated for organic anions in the statement that the 
mobility decreases at first rapidly and then more slowly 
with increase in molecular weight. For the inorganic 
ions, however, this rule does not hold; it seems, on the 
contrary, that some other influence, as in the group of the 
alkali and alkali- earth cathions, plays a part here which 


1 journ. Chem, Soc. Trans., 81 , 466 (1902). 



MOBILITY OF THE IONS . 


35 


probably must be sought for in hydration. It is 
notable that the halogens as ions, in spite of their varying 
weight and varying mobility in the form of diffusing 
neutral molecules, possess almost equal mobility. Euler 1 
offers as an explanation for this the assumption of marked 
hydration, which might equalize the difference in weight. 
Possibly this hypothesis finds support in the observation, 
that in the series of the alkalis, as well as the earths,, the 
element of strongest electro-affinity (compare p. 159), 
which may be assumed to have the least tendency to 
hydration, forms the most mobile ion. 

A very comprehensive research and summarization of 
the mobilities of all known inorganic as well as organic 
ions and the accompanying regularities, we owe to 
Bredig. 2 He found among other things that the mobility 
of the* element ions is a periodic function of the atomic 
weight; that for compound ions it essentially holds that, 
increasing the number of atoms decreases the mobility;- 
and that constitutional influences also make themselves, 
felt. 

The slowest known anion is that of the lactone of 
^-toluido-/ 9 -i-butyric acid with l a (2 5°) = 23.3; the slowest 
cathion, that of aconitine with Ik (2 5 0 ) = 17.8. 

The influence of temperature on ionic mobility has 
recently been more carefully investigated by Kohlrausch, 3 
with the result that to each ion an independent change 
of mobility can be attributed, as was to be expected in 
accordance with the additive law (p. 12). 


1 Wied. Ann., 64 , 273 (1897). 

2 Zeitschr. physik. Chem., 13 , 191 (1894). 

5 Berl. Akad. Ber., 26 , 574 (1902). 



36 THE THEORY OF ELECTROLYTIC DISSOCIATION 

These individual temperature coefficients are to be 
found under a in the above table of mobilities. They 
mean that the values for l are to be multiplied by 
(1 4- [t— i8]a) in order to obtain the values for l at ^°. It 
is of importance to note that these percentage temperature 
coefficients are smaller the larger the mobilities; the 
absolute coefficients, however, show the same order in 
the series as the mobilities, so that the ions converge in 
the direction of lower temperatures toward the same 
mobility. 

From the changes in mobility of the ions we can now 
also calculate the influence of temperature on the con- 
ductivity for other concentrations than that of extreme 
dilution, in so far as we are allowed to assume that 
essentially the ionic mobility changes and not the degree 
of dissociation, i.e., the number of ions that take part in 
the conductivity at the different temperatures. According 
to the results of investigations on this point, to be dis- 
cussed later (see pp. 135, 140), this assumption holds 
approximately for strong electrolytes and also for many 
weak ones whose heat of dissociation is small, so that 
their temperature coefficients can be calculated from the 
above values. The percentage coefficients for salts lie 
between 0.021 and 0.029, for acids in the neighborhood 
of 0.013, f° r bases near 0.020. 



EQUILIBRIA AMONG IONS. 


Starting with the conception that dissociation is to be 
considered as a chemical reaction of such a nature that 
out of the ions, the dissociation products, the undissociated 
substance is formed by chemical interaction, then we 
must look upon the law of mass action as the factor 
determining the equilibrium between the reacting ions 
and their resulting undissociated product. During the early 
days of the dissociation theory it was customary to view 
this reaction from the side of the undissociated molecule, 
the ions being formed by its decomposition. In principle, 
however, both mean the same, and it is clearer possibly 
from a chemical standpoint to consider the reaction in 
the reverse sense as we did above, and to look upon the 
ions as primary and their product, the undissociated 
compound, as secondary. Indeed, the latter seems more 
natural — though fundamentally a matter of taste — in 
so far as the presence of ions is an extremely wide-spread 
property of chemical substances. Yet not incorrectly 
perhaps and from purely historical reasons, one considers 
the inappreciably ionized compounds belonging essentially 
to organic chemistry as the normal, owing to their great 
number and the intensity of the study that has been con- 
centrated upon them. 



THE THEORY OF ELECTROLYTIC DISSOCIATION. 

If, however, we take as the normal the relations as they 
prevail with the compounds which show the greatest 
variation in the elements combining to form them, that 
is, without favoring carbon compounds, then ionic dis- 
sociation is of such- a general nature that we may place 
ionic interaction or the formation of undissociated com- 
pounds in the foreground. 

It is true, science proceeded in just the opposite way: 
the undissociated compounds were considered the normal 
ones. The formation of ions was formerly unknown and 
in a certain sense did not take place until the introduction 
of the theory, because not until then was it a conscious 
change. But however that may be, the compounds which 
are very little ionized represent such whose components 
are held together by. exceptionally strong forces of atomic 
affinity, while the chemical relationship between the 
components (ions) of the strongly dissociated substances 
must be considerably less, in order to make possible for 
them the independent existence. 

A thing difficult of conception also lies in the assumption 
that a reaction is to arise out of an undissociated substance 
without that substance interacting with other substances; 
this in fact becomes inconceivable when one assumes the 
hypothesis, to be discussed later, that all reactions are 
dependent on the presence of ions, and that , their velocity 
is directly determined by the concentration of the ions 
necessary for the reaction. Suppose we picture to our- 
selves a molecule, capable of ionization, in an entirely 
undissociated state, then a dissociation into ions cannot 
take place at all, because according to our assumption 
the undissociated substance was to have no ions. We 



EQUILIBRIA AMONG IONS. 


39 


can readily conceive of the reverse, for when ions are 
once present, undissociated substances form by their 
reaction. For this it would be necessary, of course, to 
assume that no chemical element could exist in the state 
of an absolutely electricity-free non-ion. However, we 
shall not here continue these speculations, for they are 
of no consequence as far as the numerical laws of dis- 
sociation are concerned. 

For the dissociation of substances such as ammonium 
chloride, phosphorus pentachloride, and others, which in 
the gaseous state split up into simpler components, the law 
of mass action has shown that the product of the con- 
centrations (partial pressures) of the reacting constituents 
is proportional to the product of the concentration of the 
substances produced by the reaction. Now precisely 
the same mathematical relation must also hold for dis- 
sociation into ions and the reaction of ions to form un- 
dissociated molecules. And it was the great service of 
Ostwald to have recognized this law and confirmed it for 
a much wider range than that for which it had been estab- 
lished for the then known gas dissociations. This law reg- 
ulates the concentration of the ions and undissociated 
molecules with varying total concentration of the solution. 
In other words, it places us in position to derive from the 
degree of dissociation of an electrolyte at one concentration 
the degree of dissociation at any other desired concentra- 
tion. If, for instance, for the concentration c, the degree 
of dissociation, or the part per mole split up, is a , then the 
total concentration of each ion is a -c, and the concentra- 
tion of the undissociated remainder is (i —a) c. Intro- 
ducing this value into the law of mass action, we get for 



40 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


the product of the concentrations of the reacting sub- 
stances (the ions), for the simplest case of a binary 
electrolyte in which two ions unite to form an undissociated 
molecule, 

(a • c) • (a - c). 


This product is to be proportional to the concentration 
of the undissociated substance, namely (i — a) Hence 
for a binary electrolyte the expression for the law of 
mass action is 


a 2 'C 2 =^k -( i —a) c, 

if k indicates the proportionality constant which is 
characteristic (at a definite temperature) for this reaction. 
Now having found the degree of dissociation a for the 
concentration c, by means of one of the above methods, 
i.e., conductivity or osmotic pressure measurements 
(freezing-point, boiling-point, etc.), we can calculate 
by the given formula the “ dissociation constant ” k and 
are then in position to determine the degree of dissociation 
for other concentrations (c values) with the aid of the 
transformed equation 


a 2 k 
i —a c 


\ 3 ) 


or 


— £+\/£ 2 +4 kc 

a = . 


2C 



EQUILIBRIA AMONG IONS. 4 * 

or neglecting k 2 and k as compared with \ which is 
permissible for small values of k, we have approximately 

Ik 

a = y!- ( 3 a ) 

The testing of this important relation, the so-called 
dilution law of electrolytes discovered by Ostwald , 1 was 
first undertaken by van’t Hoff and Reicher 2 on a series 
of acids and resulted in an excellent confirmation. The 
authors close their discussion with these words: “Not 
a single case of ordinary dissociation has been tested 
within such wide limits.” Some of the figures are given 
in the following tables: 


Acetic Acid: £ = i.78X io ~ 5 (14 0 ). 


_ J 

c 

0-994 

2-02 

15-9 

1S.1 

1500 

3010 74S0 

15000 

100 a from conductivity A . . 

0.40 

0.6l 

1.66 

1.78 

14-7 

20.530.I 

4O.8 

100a from k calculated 

p.42 

0.60 

1.67 

1.78 

15 -° 

20.2 3O.5 

40.1 


Monochloracetic Acid: k = i .$ 8 Xio - 3 (14 0 ). 


_ X __ 

c 

20 

205 40S 

2060 4080 

IOIOO 

20700 

100 a from A found 

16.6 

42.3 54-7 

80.6 88.1 

94.8 

96.3 

1 00a from k calculated. . . . 

16.3 

43 -° 54-3 

80.1 S8.0 

94-4 

97.1 


Ostwald about the same time, in amassing his extended 
observations on the organic carboxylic acids, used a 


1 Zeitschr. physik. Chem., 2 , 36 (1S8S). 

2 Ibid., 2 , 777 (18SS). 




42 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


method, which since then has remained the customary 
one, so that usually the law of mass action is tested not by 
comparing observed and calculated degrees of dissociation, 
but by testing whether the expression for the characteristic 
dissociation constant. 


gives values independent of the dilution. Since in the 
majority of cases a , the degree of dissociation, is obtained 

with the aid of conductivity, i.e., from it is practical to 


insert this expression directly into the formula, giving it 
the form 


or 



A 2 -c 

A 0 (Ao-A) 


= k, 


( 4 ) 


or finally, introducing the specific conductivity k=A-c: 

* _* 

A g (A 0 -c-k) 


For the relatively frequent case of very weakly dis- 
sociated electrolytes, in w T hich the degree of dissociation 
a is only a small fraction (say 1% or less) of the total 
concentration, the general formula can be conveniently 



EQUILIBRIA AMONG IONS . 


43 


simplified by writing i — a = i, by reason of the smallness 
of a , when it becomes 


a 2 -c = k or 


A 2 c 

A 0 2 


( 5 ) 


From the latter equation a very simple law for the 
variation of conductivity may be derived. Since J 0 for 
one and the same electrolyte is constant, being independent 

r , . , . Const, 

of the concentration, we have simply = — - — or A 

inversely proportional to Vc. Finally, the equivalent 
conductivity being A = wherein k is the specific con- 
ductivity, we can, by substituting in the last equation, 
so formulate the relation between the specific conductivity 
k and the concentration of the electrolyte that 

a : 2 — Const . - c or jc = Const . * V 7 c, 


which means, in other words, that the conductivity of a 
solution with varying concentration is proportional to 
the root of this concentration. For example, diluting a 
solution four times reduces the conductivity only one 
half, or diluting ten times reduces it only 3.16 times. 
This is shown by the following small table for acetic 
acid, taken from measurements of Kohlrausch: 


. Acetic Acid (iS°). 


C— 1.0 

O 5 

0.1 

0.05 

0.0 1 

0.005 

0.00 1 

0.0005 

0.0001 

K = I* 3 2 
.4=1.32 

1.005 
2 0 : 

0.46 
,1 . 60 

0.324 

6. 4 3 

0.143 

14 - 3 . 

0. 100 

20.0 

0.041 

41.0 

0.0285 

57 -o 

6.0107 

107.0 


44 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


The following tabulated figures are taken from the 
previously mentioned measurements of Ostwald, given 
for the greater part in Zeitschr. physik. Chem., 3 , 170, 
241, 369 (1889), as well as from those of Bredig. 1 Here 

IQ ° : - expresses the degree of dissociation in percentages. 

- 1 0 


I 

Acetic Acid : 

4 o =3 SS 

Monochloracetic Acid : 
Aq — 386 

Dichloracetic Acid: 
A 0 = 3S5 

v— — 
c 

A 

iood 

- 4 o 

io 5 fe 

A 

TOO A 

Ao 

io 5 k 

A 

. 

100 4 

^0 

io 5 & 

16 

6 -5 

1-67 

i- 79 

56.6 

14.6 

J 55 

269.8 

— 

— 

32 

9.2 

2.38 

1.S2 

77.2 

19.9 

I 55 

70. 2 

5170 

64 

12.9 

3-33 

1.79 

IO3.2 

26.7 

152 

309-9 

80.5 

520° 

I2S 

18. 1 

4.6S 

i -79 

136. 1 

35 - 2 

150 

338-4 

‘88.0 

5040 

256 

25-4 

6.56 

1 .So 

174.8 

45.2 

146 

359-2 

93-4 

5 i6 ° 

5 12 

34-3 

9.14 

1.80 

219.4 

q6.S 

146 

375-4 

97.6 

— 

1024 

49-0 

12.66 

i -77 

265.7 

6S.7 

147 

383-8 

99-7 



Ammonia: Aq=^ 253 

Methylamine. J 0 = 240 

Piperidine: A 0 

= 216 

T 

v= — 
c 

A 

ioo 4 

Aq 

io 5 k 

A 

100A 

4 ) 

io 5 k 

A 

100A 

Ao 

IQSJfe 

8 

3-4 

i -35 

2 -3 

I 5 - 1 

6 -3 

52 

23.O 

10.6 

157 

16 

4.8 

1.88 

2 -3 

21.0 

8.7 

5 2 

3 2 '3 

14.9 

163 

3 2 

6.7 

2.65 

2 -3 

28.9 

12.0 

5 i 

44.2 

20.3 

162 

64 

9-5 

3- 76 

2 -3 

39-3 

16.3 

50 

59-2 

27.2 

159 

128 

13-5 

5-33 

2 -3 

53 -o 

22.0 

49 

77.8 

35-8 

156 

256 

18.2 

7-54 

2.4 

70.0 

29.I 

47 

99-7 

45-9 

152 


That the measurements of the degree of dissociation 
from determinations of the freezing-points lead to the 
same results is shown by the following series of observa- 


1 Zeitschr. physik, Chem., 13 ? 289 (1894), 


EQUILIBRIA AMONG IONS . 


45 


tions: A indicates the depression of the freezing-point, 
1.85 the depression in water of each mole of undissociated 
substance. 


Tartaric Acid. 
(Abegg, 1896.) 


c 

1 

c 

I - S A 

a—i — 1 ! 

H 

a 

C 

0.00516 

2.45° 

I.32 

0-32 ; 

°- 35 

0.0103 

2.29 

1.24 

0 

to 

4— 

0.26 

0.0154 

2 . 24 

1. 21 

0.21 

0. 22 

0 . 0204 

2.23 

I.205 

0.205 

0. 20 

0.0254 

| 2.18 

I. IS 

0 . iS 

0. iS 

0 . 0303 

2.15 

I .16 

0. 16 

0. 16 

°-°353 

1 

2 -OS 

1. 12 j 

0.12 

0.15 


The extent of the observations confirming the dilution 
law — in other words, showing the validity of the law of 
mass action — -may be seen by a glance at the comprehensive 
tables to be found very systematically arranged in the 
excellent book of Kohlrausch and Holbom, 1 “ Leitver- 
mogen der Elektrolvte,” pp. 176 to 194. 

Of this material the greater part relates to weak organic 
acids and is taken from the researches of Ostwald and 
his pupils, 2 among whom Walden is to be especially 
mentioned. A smaller part consists of the measurements 
of Bredig 3 on bases, among which especially the weak 
amine bases confirm the dilution law. 


1 Teubner, .Leipzig, 1898. 

3 The complete literature is to be found in the mentioned work of 
Kohlrausch and Hqlborn. 

* Zeitschr. physik. Chem., 13, 2S9 ( 1894 ). 



THE DISSOCIATION CONSTANT. 


This great mass of material naturally offered not only 
a confirmation of the mathematical formulation of the 
relation between the degree of dissociation and the 
concentration of the electrolyte, but also enabled us to 
gain important chemical knowledge from the measure 
of the dissociation constant. This constant is indeed an 
expression of the chemical nature of substances, in that 
it gives a numerical measure of the tendency to split into 
ions. If we do not apply the above form of the dissocia- 
tion constant given by Ostwald, but rather its reciprocal 

value p then this would constitute an analogous numerical 

expression for what we have previously termed the atomic 
affinity, which exists between ions and tends to produce 
undissociated molecules out of them. 

The physical significance of the constant k can also be ex- 
pressed, with Ostwald, as indicating one half of that concen- 
tration at which the various electrolytes possess exactly the 
degree of dissociation equal to J. For example, taking 
for comparison the constant of acetic acid (0.000018), 
of monochloracetic acid (0.00155), °f dichloracetic acid 
(0.051), and also let us say of malonic acid (0.00158) 
and maleic acid (o.oi2) ? it means that these acids are 

46 



THE DISSOCIATION CONSTANT. 


47 


dissociated 50% in solutions which for acetic acid have 
the normal concentration 0.000036, for monochloracetic 
acid 0.0031* for dichloracetic acid 0.12, for malonic 
acid 0.00316, and for maleic acid 0.024. The definition 
for \/ k taken from the formula (3 a), p. 41, is pos- 
sibly clearer, \/k being the ionic concentration present 
in the i-normal solution of the electrolyte. Since the 
action of acids is determined by the concentration of 
the hydrogen ion, that of bases by the hydroxyl ion, it is 
easy to see the great value of knowing this dissociation 
constant in comparing chemical nature, and it was to be 
expected from the very first that this characteristic 
constant should bear a marked relationship to the chemical 
constitution of these substances. This has in fact been 
found to be true, and as it is our desire to trace at least 
the bolder outlines of this relationship between chemical 
nature and the dissociation constant, we shall bring 
together in 'the following tables the dissociation constants 
of some interesting acids. 

Substitution or CH 3 . 


Formic add, HCOOH £=127.0 Xio— 5 

Acetic acid, CH 3 COOH 1.8 Xio -5 

Propionic add, CoH 5 COOH 1.3 Xio~ 5 

Butyric acid, CaHyCOOH 1.5 Xio~ 5 

Isobutyric add 1.45X10- 5 

Valeric add, QH^COOH 1.6 Xio~ 5 

Caproic add, C5H11COOH. 1.45X10 -5 


While the first substitutions without doubt produce a 
weakening of the acid, the very first as much as seventy 
times, the subsequent ones are occasionally accompanied 
by . a slight strengthening. 



48 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


Substitution of Halogens. 


Acetic acid 

Chloracetic acid 

Dichloracetic acid 

Trichloracetic acid 

Bromacetic acid 

Cyanacetic acid. . . 

Sulpho-cyanacetic acid 

*55 

5 100 

138 

— 370 

Lactic acid 

Trichlorlactic acid 


Propionic acid 

/?-Iodpropionic acid 

9.0X10- 5 

Benzoic acid 

w-FIuorbenzoIc acid 



Here one sees that all these substitutions bring about 
a very marked strengthening, and again that with several 
successive ones — as in general — the first step is the most 
effective; furthermore, that the proximity of the substitut- 
ing groups is of great influence, as will later be pointed 
out more fully. 


Substitution of OH. 

Acetic acid, CH 3 COOH io 5 X k= i . 8 . 

Glycollic acid, CHjOHCOOH 15.0 

Propionic acid, CH 3 CH 2 COOH 1.3 

Lactic acid, CHjCHtOHJCOOH . 14.0 

^-Oxypropionic acid, CH^OH) CH 2 COOH . 3.1 


Benzoic acid, C 6 H 5 COOH 6.0 

Salicylic acid, C 6 H 4 (OH)COOH (1:2).... 102.0 

•w-Oxybenzoic acid, C 6 H 4 (OH)COOH (1:3) 8.7 

^-Oxybenzoic acid, C e H 4 (OH) COOH (1:4) 2.9 



THE DISSOCIATION CONSTANT . 


49 


The nearer to the COOH group the OH is introduced, 
the more it increases the dissociation of acids; the same 
is true of N 0 2 and COOH: 


Substitution of NO*. 

Benzoic acid io 5 X&= 6.0 

<?-Nitrobenzoic acid 620 . o 

w-JNTitrobenzoic acid 35 .0 

^-Nitrobenzoic acid 40.0 

Phenol £=1.3 Xio~ 10 

n-Nitrophenol 4.2 Xio~ s 

2,6-Dinitrophenol 1.7 Xio" 4 

Trinitrophenol 1.64X10— 1 t 1 ) 

Salicylic acid io 3 X£ = 1.02 

n-Nitrosalicylic acid 115 .7 

^-Nitrosalicylic acid 9.0 

Substitution of COOH. 

Acetic acid, CH 3 COOH 10 s X k = 1 . 8 

Malonic acid, COOHCH 2 COOH 158. o 

Propionic acid, C 2 H 5 COOH : . 1.3 

Succinic acid, COOHC 2 H 4 COOH 6.6 

Benzoic acid, C e H 5 COOH 6.0 

0-PhthaIic acid, COOHC 6 H 4 COOH 121 . o 

w-Phtfcalic acid, COOHC e H 4 COOH. ... 29.0 


Substitution of NH 2 

exceptionally weakens the acid character, so that on the 
one hand the very strong sulpho-acids, whose constants, 
for reasons to be given later (see p. 121), lie beyond those 
capable of measurement, are brought by substitution 
within the scope of those measurable, while on the other 
hand the acids of the average strength of the above are 
decidedly weakened. A constant for these is not to be 

■ 1 Rothmund and Drucker, Zeitscbr. physik. Chem., 46, 827 (1903), 



5o THE THEORY OF ELECTROLYTIC DISSOCIATION. 

obtained directly by conductivity measurements, on ac- 
count of their capacity for amphoteric (acid and basic) 
dissociation; however, B redig and Winkelblech 1 showed 
how both the acid and basic dissociation constants may 
be obtained (see table, p. 55). 


Influence of the Position of the Substituting Groups. 


In addition to the preceding characteristic examples, 
the series of dicarboxylic acids may be given: 


Oxalic acid, COOH.COOH 

Malonic acid, COOH.CH 2 .COOH. . 
Succinic acid, COOH.C 2 H 4 .COOH. . 
Glutaric acid, COOH.C 3 H G .COOH . . 
Adipic acid, COOH.C 4 H S .COOH . . . 
Pimelic acid, COOH.C 5 H l0 .COOH . . 
Suberic acid, COOH.C 6 H 12 .COOH . . 
Sebacic acid, COOH.C s H 1G -COOH.. 


io 5 Xj&= about 10000 

158 

6.65 

4-75 

3-7 

3-6 

2.6 

2 -3 


Methylmalonic acid, 

COOH.CH(CH 3 ).COOH 

Pyrotartaric acid, 

COOH-QHaCCEy. CO OH 

„ . COOH.CH 

Fumanc acid, r 3v . 

CH.COOH... io 3 X&= 

, CH.COOH 

Male* aad, g H-COO jj 


87 

8.6 

o-93 

11.70 


In these instances the essential relation coming into 
play is the distance apart of the two COOH groups: 
the more carbon atoms between them the weaker the acid 
becomes — here again the first steps are decidedly the most 
effective. Upon attaining a certain separation a further 
increase of the distance makes less and less impression. 


1 Zeitschr. physik, Chem., 36 , 546 (1901). 


THE DISSOCIATION CONSTANT. 


5 * 


Comparing fumaric and maleic acids, the action of the 
proximity of the two carboxyl groups is especially apparent. 

In the case of organic compounds one condition for the 
production of H' ions is evidently the direct union of H 
and O; for that reason the alcohols show a distinct, 
even if extremely slight, acid function (the alcoholates). 
Another important condition is the proximity of carbonyl 
groups, which, for example, in the case of malonic-acid 
ester, acetic-acid ester, and also in acetylacetone makes 
the hydrogens in the neighborhood of the CH 2 groups 
capable of dissociation and salt formation. 1 The carboxyl 
compounds undoubtedly owe their marked acid property 
to the combination of both conditions. In the repre- 
sentatives of the first two groups of compounds the 
dissociation is scarcely detectable by physical means; 
therefore the decomposition of their salts by water (see 
Hydrolysis, p. 76) is almost complete. 

In the case of bases, all substituting groups have just 
the reverse action of that on acids; the halogens, the 
carboxyl group, and the N 0 2 group have an especially 
weakening effect on the basic character. A detailed in- 
vestigation of Bredig 2 gives a fuller account of this. 
Because of the great influence of constitution a quan- 
titative determination of the eff ect of substitution has up 
to the present not been possible. 

Another phenomenon deserves special mention, which 
was likewise first noted and made clear by Ostwald, 
nam ely, the dissociation of the dibasic organic acids. 
As these contain two CO OH groups, there is a possibility 

1 Ehrenfeld, Zeitschr. f. Elektrochem., 9 , 335 (1903). 

2 Zeitschr. physik. Chem., 13 , 289 (1894). 



52 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


of a hydrogen ion dissociation taking place at both of 
these, the more so as we have just seen that the presence 
of a second carboxyl group in the molecule markedly 
increases the dissociating tendency of the first. Now, 
strange to say, a calculation of the constant by equation 

(3), — — which holds only for binary electrolytes 

(those splitting into two ions), shows that this is not the 
case. For if both carboxyl groups split off H* ions, such 
an acid would have ternary dissociation, that is, would 
have to obey another dissociation formula. One is forced, 
therefore, to the view that the dissociation at the second 
carboxyl group takes place with considerably greater diffi- 
culty than at the first, so that an influence of such a nature 
must be present that the first step of the dissociation 
prevents the second from taking place. That a second 
stage sets in at all, can be recognized by the fact that in 
general the binary dissociation formula ceases to apply 
when the degree of dissociation of the first stage has 
reached about J, for here the constancy of the expression 
cx?-c 

- — — ceases (compare table on opposite page). The 

physical significance of this phenomenon Ostwald finds 
in that the presence of a negative charge on a univalent 
acid anion makes more difficult the placing upon it of a 
second ionic charge, for the electrostatic reason that like 
charges repel one another. This second ionic charge 
would be necessitated by a dissociation at the second 
carboxyl group. One consequence of this view may be 
empirically tested: the appearance of the second stage of 
dissociation would have to be influenced by the relative 



Second Dissociation of Dibasic Acids. 
(Ostwald, 1889.) 


THE DISSOCIATION CONSTANT. 



Fumaric acid, COOH.C 2 H 2 COOIL 



54 the theory of electrolytic dissociation . 


position of the two carboxyl groups in the molecule, for 
it is evident that the electrostatic interaction must be 
greater the nearer the negatively charged carboxyls 
are to each other. This surmise is fully confirmed. In 
dibasic acids whose constitution shows a close proximity 
of the two carboxyls, the second dissociation sets in 
with considerably greater difficulty than in acids where 
the carboxyls are farther apart. In the enumeration on 
p. 53 are given the constants k of different dibasic 
acids calculated according to the dilution law for binary 
dissociation. At the dilution marked f the binary 
constants increase, showing the beginning of the second, 
the ternary stage of dissociation. Under a the two 
degrees of dissociation are given, between which the 
ternary dissociation begins. The most marked evidence 
of the influence of the proximity of the carboxyls is given 
by fumaric and maleic acids and the phthalic acids with 
the adjacent position of the two CO OH groups. 

As is to be seen, w-hen the two COOH groups are near 
together, in spite of far-reaching primary dissociation, the 
secondary does not set in until very late (a-nitrophthalic 
acid) or not at all (maleic acid), while in most cases it 
begins with a equal to about 0.5. 

Of especial interest are the extremely weak electrolytes, 
previously mentioned, a list of which is appended here 1 
(for 25°) : 

1 From Walker and Cormack, Journ. Chem. Soc., 77 , 5 (1900); Zeit- 
schr. physik. Chem., 22 , 137 (1900). — Walker, ibid., 4 , 332 (1889), and 
32 , 137 (1900). — B redig, ibid., 13 , 322 (1894). — Winkelblech, ibid., 36 , 
587 (1901). — Lowenherz, ibid., 25 , 385 (1898). — Morse, ibid., 41 , 709 
(1902). — Bader, ibid., 6, 289 (1890). — Walker and Wood, Proc. Chem. 
Soc., 19 , 67 (1903). 



THE DISSOCIATION 

CONSTANT . 

55 

Extremely Weak Electrolytes. 


Acids: 



Meta-arsenious acid, H*, AsO/ 

. 2.1 Xio~ s 


w-Amidobenzoic acid, H', 



C 6 H 4 (NH>)COO' 

9.6 X io~ 6 


Carbonic acid, H*, HCO s ' 

3.04X IO“ 7 


^-Nitrophenol, H*, C G H 4 (NO.AO' 

1.2 X io~ 7 


Hydrogen sulphide, H*, SH' 

5.7 Xio~ s 


Boric add, H*, H^BO/- 

1.7 Xio-® 


Hydrocyanic acid, H', CN' 

1.3 Xio- 9 


Alanine, H-, CJS 5 (NEQCOO' 

9.0 Xio -10 


Phenol, H*, C 6 H s O' 

1.3 Xio- 10 


Water (25 0 ), H*, OH' 

. 1 . 2 X io~ 14 (ionic product) 

Cacodylic acid, H', (CH 3 )^AsOO' 

. 4.2 Xio- 7 


Bases: 



^-Cumidine, OH', C^CH^H, . . . 

1.7 Xio- 9 


^-Toluidine, OH', C-H-NH-f 

1.6 Xio- 9 


Aniline, OH', C G H 5 NH 3 * 

4.9 Xio- 10 


w-Amidobenzoic add, OH', 



C 6 H 4 (COOH)NH 3 * 

1.9 X io~ u 


m-Nitxaniline, OH', C 0 H,(NO 2 )NH 3 - . . 

4.0 X io“ 12 


Alanine, OH', CJH 5 (COOH)NH 3 ' 

3.8 Xio- 13 


Tbiazole, OH', CftSNH* 

3.3 Xio- 12 


GlycocoU, OH', CH.(COJEI)NH 3 

2.9 Xio -12 


Asparagine. OH', 



C 2 H 3 (C 0 2 H)(C 0 NH 2 )NH 3 ' 

1.3 Xio- 13 


^-Nitraniline, OH', C 6 H 4 (NO)oNH 3 -. . . 

1.0 Xio -12 


Thiohydantoin, OH', C 3 H 5 N 2 SO* 

9.5 Xio- 13 


Aspartic acid, OH', CoH 3 (C 0 2 H) 2 NH 3 ' 

8.7 Xio^ 13 


Betaine, OH', CH^CO^NCCH,)^. . . 

7.6 Xio— 13 


Acetoxime, OH', (CH 3 ) 2 CNHOH* . 6. iX 10— 13 (25°); 1.8 Xio - 

13 (40°) 

Urea, OH', CONJEJ,.* 1.5X10- 14 (25 0 ); 3.7X10- 

14 (40°) 

o-Nitraniline, OH', C 6 H 4 (N0 2 )NH3’ . . . . 

1.0 Xio -14 


Water (25°), OH', H’ 

1.2 X io- 14 (ionic product) 

Acetamide, OH', CH 3 CONH 3 * 

3.oXicr- 15 (25°); 



3.3X10- 

14 (40°) 

Propionitrile, OH', CJH 5 CNH‘ 

1.8 Xio- 15 


Thio-urea, OH', CSN*H s * 

1.1 Xio- 15 


Cacodylic add, OH', (GHQsAsO' . 2.5X 

IO— 13 (23°); 3.8X10 

- 13 (o°) 

Salts: 



Mercuric chloride, HgCL, 

i X IO- 14 


Mercuric bromide, HgBr 2 

2 Xio- 18 


Mercuric iodide, Hgl 2 

iXio-« 




the theory op electrolytic dissociation . 

Among these, the dissociation of pure water into H* 
and OH 7 ions is of particular importance. This constant 
at room temperature is equal to about io~ 14 , i.e., the 
product of the hydrogen and hydroxyl ion concentrations 
has the above value, or in pure water each kind of ion is 
present in the concentration io -7 . In other words, pure 
water is one ten-millionth normal with reference to the 
hydrogen and hydroxyl ions . 1 

This value has been arrived at in four entirely inde- 
pendent ways, and the different results show excellent agree- 
ment. Kohlrausch and Heydweiller 2 3 determined the con- 
ductivity of water purified with extreme care, after they 
had discovered that the conductivity of the common dis- 
tilled water is for the most part due to such substances as 
carbonic acid, ammonium salts, glass, etc., dissolved from 
the atmosphere and the walls of the vessel. By repeated 
distillation in vacuo in specially prepared vessels of most 
sparingly soluble glass, they succeeded in obtaining a 
conductivity, 

k=o.04Xio"~ 6 (i8°), 

which in conjunction with the mobility of the H* and OH' 
ions gives the named ionic concentration; since i mole 

1 The constant io~ 14 is in the true sense not a dissociation constant, 

but merely represents the ionic product, that is, (IT) - (OH') or k- 1 H 2 0 ) ; 
since, however, on the one hand the concentration of the BL 2 0 molecules 
in water is unknown (on account of polymerization), and on the other 
hand practically does not vary to any extent in dilute solutions, it is 
to no purpose to introduce for the ionic equilibria in which water takes 
part any other constant than the ionic product, namely, k ■ (H 2 Q), which 
in the future shall be designated by k or the ‘‘water constant.” 

3 Wied. Ann., 53 , 209; Zeitschr. physik. Chem., 14 , 317 (1S94). 



THE DISSOCIATION CONSTANT. 


57 


H-+OH' ions in i c.c. would produce (see p. 33) the 

conductivity 318 + 174 = 492, therefore there are only 

o.o4Xio“ 6 , , 

mole 10ns m 1 c.c. = o.8Xio _/ mole per 

492 

liter. 

Ostwald followed a second method requiring much 
less precision. He measured the electromotive force of 
two hydrogen electrodes opposed to each other, the one 
dipping into an acid, the other into an alkali, of known 
H“ and OH' concentration respectively. This galvanic 
combination can be looked upon as a concentration chain 
of hydrogen ions, whose force, according to Nemst’s 
theory, serves to determine the H' ion concentration in 
the hydroxide solution employed, and in consequence 
permits the calculation of the product of H* and OH' ion 
concentrations. This product represents the dissociation 
constant of water. The result was the same as above. 
(We shall learn later how in such cases as this, where the 
concentration of the two ions is very different, the law 
of mass action is applied.) 

A third way, that led to the same result, was the meas- 
urement of the rate of saponification of esters by water, 
as carried out by Wijs according to a theory of van’t 
Hoff. 1 

Finally, Shields, 2 at the suggestion of Arrhenius, 
studied the hydrolysis of salts, which, as we shall have 
to consider later, allows the calculation of the dissociation 
constant of water. This constant proves to be the same 
as given above. 

1 Zeitschr. physik. Chem., 12 , 514 (1893). 

2 Ibid., 12 , 167 (1893). 



5 & THE THEORY OF ELECTROLYTIC DISSOCIATION . 

Let us now discuss one of the most interesting 
conclusions from this extremely small water disso- 
ciation, namely, the process of neutralization of acids 
by bases. The slight dissociation of water is nothing 
more than the expression of the fact that H* and OH' 
ions possess a very strong affinity for each other, so that 
the extent to which they unite to form undissociated 
waterjs.scf complete that only the lepeatedly mentioned 
very small number of H* and OH' ions is left. If there- 
fore 'these* ions meet in any solution in higher concentra- 
tions, they, cannot be in equilibrium with one another, 
but must continue to unite to form undissociated water 
until the product of their concentrations remaining 
has reached the value io“ 14 . Therefore upon mixing 
equivalent solutions of H’ ions (acids) and OH' ions 
(alkalis) the union of these ions to form undissociated 
water will set in above all other things, aside from any 
further reactions. Whether the anions of the acid and 
the cathions of the alkali undergo further chemical action 
with one another is of course a question by itself. For 
ordinary cases this question is, however, to be answered 
in the negative, since, as alluded to above, salts, which 
would have to be formed by the combination of these 
two kinds of ions, are for the most part strongly dis- 
sociated, i.e., consist of ions, so that these ions find no 
occasion to form any marked quantities of undissociated 
salts. We see then that the essential change taking place 
upon mixing acids and bases is the formation of un- 
dissociated water by the H* and OH' ions. One con- 
clusion from this conception has been known for a very 
long time, ever since the investigations of the Russian 



THE DISSOCIATION CONSTANT. 59 

thermochemist Hess, 1 who made the most startling 
discover} 7 , and one at that time inconceivable, that the 
heat effect of neutralization of dissolved acids and bases 
in equivalent amounts always gave the same value, 13700 
cal. per gram-equivalent. On the basis of the dissociation 
theory this fact could have been predicted, for in all these 



in the ionic sense becomes 


K-+OH'+H-+Cl / =K-+Cl'+H 2 0 , 

and leaving out the unchanged substances on the right 
and left sides, the ions K* and Cl', we arrive, as you see, 
at the above simple equation H* + 0 H / =H 2 0 for the 
process of neutralization. A test of the question whether 
this heat effect really has the significance of a heat of 
dissociation of water, as the simple neutralization equation 
represents it, has been possible in another way, namely, 
by the investigation of the variation of the water dis- 
sociation a with the temperature T. Thermodynamical 
considerations give the following mathematical relation 
of a and T to the heat of dissociation W (p. 137): 

1 dec W 
a"df~ 2 RT 2 


1 Ostwald’s Illiussiker, Nr. 9, 1S39-1S42. 



60 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

Kohlrausch and Heydweiller experimentally tested 
this variation of the dissociation of water with the tem- 
perature by means of the conductivity of pure water, and 
found that the heat of dissociation, calculated by the 
van’t Hoff equation given above, gave results in complete 
agreement with the heat of neutralization as determined 
by Hess and Thomsen. (In reality the reverse, which 
in principle means the same thing, was done; that is 
to say, the variation of the degree of dissociation with 
the temperature was calculated by van’t Hoff’s equation, 
on the assumption that the heat of neutralization really 
represents the heat of dissociation of water.) 

Another thermochemical result of Hess is explained 
very nicely by the dissociation theory, namely, the 
thermoneutrality of salt solutions, or the fact that mod- 
erately dilute salt solutions when mixed together give no 
heat effect — in other words, show no signs of reaction. 
This in spite of the fact that, according to our old views 
in such a process of mixing, at least a partial mutual 
decomposition of both salts with the formation of new 
salts ought to take place. According to the dissociation 
theory, however, the ions are for the most part free before 
and after the mixing, and therefore no reaction takes 
place; for it is extremely improbable that in all these 
various cases the heat effects of the reactions taking place 
would just compensate each other, making the total 
effect equal to zero. 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 


The dilution formula for binary electrolytes in the form 
oc 2 c 

given above, ^—^== Const., evidently holds only on the 

assumption that both ions of the electrolyte are present 
in equivalent amounts, which is necessarily correct as 
long as no second electrolyte is present in solution at the 
same time. It is frequently the case, however, that in a 
solution two electrolytes are present which have one of 
their ions in common, as, for instance, two acids or two 
bases, each of which forms H* and OH' ions respectively; 
or a salt and an acid, as sodium acetate and acetic acid; 
or a salt and a base, as ammonium chloride and ammonia. 
In the last case the ammonium cathion, in the preceding 
the acetate anion, are the ions in common. Now in the 
same manner as the gaseous dissociation of PCI5, for 
example, is affected by the addition of chlorine or PCI3, 
just so the addition of a “ like-ioned ” electrolyte must 
influence the dissociation of a co-solute, and indeed the 
law of mass action gives here, as above, the quantitative 
relations. Suppose the two electrolytes to be binary, 
and and k 2 the constants which regulate the equi- 
librium between the ions and the undissociated portion 

6 * 



62 THE theory of electrolytic dissociation . 

of the electrolytes i and 2 ; then in the common solution 
the conditions of the equilibrium for each electrolyte. 

Product of the ionic concentrations 
Concentration of tne undissociated portion * 

must be fulfilled. 

If we indicate by c\ and respectively, the total con- 
centrations of the two electrolytes and by cti and 
their degrees of dissociation, then the concentrations of 
the ions not mutual, that is, those which are produced 
by one of the electrolytes alone, are equal to aq-Ci and 
a 2 -c 2 , respectively; while the concentration of the 
mutual ions is made up of those formed from each of 
the two electrolytes, and hence is represented by the 
expression 

(X 1 -C\ +« 2 ' c 2 - 


Therefore in the common solution we have the following 
relations: 


«i-Ci-(n:i-Ci + a: 2 -c 2 ) 
1— (1— ax) -c x 

and 

, <*2-g2-(«i-ci + g 2 -C2) 

2_ (i-a 2 )-c 2 


~^i^ici+a 2 c 2 ), 

— — — (otiCi +a 2 c 2 ). 

1 — a 2 J 


( 6 ) 


The most important application of this double formula 
was made by Arrhenius in his theory of isohydric solutions , 1 
which says that upon mixing the solutions of two electro- 
lytes having a common ion no change in the degree of 


1 Zeitschr. physik. Chem., % 284 (188S). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 63 

dissociation of either takes place, if the concentrations 
of the common ion are the same in both solutions before 
mixing. 

The correctness of this statement becomes evident 
at once when we consider the following example. Assum- 
ing that we have a weak acid HA at any concentration c, 
then we may write 

(H')(AQ 
kl ~ (HA) • 

Now if we dilute this acid HA by the addition of such 
a solution of a second acid in which the concentration 
of the H’ ions is just as great as in HA, or, as Arrhenius 
puts it, an acid of isohydric concentration, it has no 
influence on the concentration of the H’ ions, while the 
anions A' as well as the undissociated molecules HA 
are both diluted to the same extent. The effect, therefore, 
of this dilution disappears in the expression 

rTn (AO , 

or in such a case the condition of dissociation remains 
unchanged and independent of the mixing ratio ; for with 
unchanged H* concentration the H‘ , A' equilibrium 
requires that the concentration ratio of anion to undis- 
sociated portion be kept the same as in the pure solution 
of the acid. 

It is clear that we can look upon any mixture of two 
acids as composed of such quantities of each pure acid 
solution as are isohydric with one another. These 



64 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


isohydric concentrations may be arrived at through the 
following consideration. In the mixture suppose C\ to be 
the concentration of the acid HAi, c 2 that of HA 2 , (H") 
the total concentration of H* ions, determined by con- 
ductivity, catalysis, inversion, or in some other way, and 
finally a\ and a 2 the respective degrees of dissociation; 
then 


£1 — 


(H , )(A / x) . aid 

(HAi) ^ ; (i-ai)ci 


-(HO 


1 —<X\ 


k%= 


(HQ (A' 2 ) 
(HA 2 ) 


-(HO 


a *2 _ m .N «2 

(1— a 2 )c 2 ^ J i—a 2 


from which 




(HO 

h ' 


The sought-for concentrations x± and x 2 of the pure 
solutions, whose degrees of dissociation are also a± and 
a 2 , are given by the relations: 


V <*l 2x l ' _ h 1 -«1 
k \ — , X\ — ki 


1 -ai 




2 > 


k 2 — 


a^x 2 
1 — ol 2 


x 2 — h 


1 — 0 : 2 , 
: & 2 2 3 


or by substituting for a\ and a 2 the values for ki and 
k 2 , respectively, found above: 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 65 


or by replacing a with k : 





-(HO 



and 


x 2 = (H*) 


(H*) 2 
*2 * 


in place of which, for small k values, the approximation 
usually suffices: 


( H ‘) 2 1 

Xi = 7 — and 



X'l _ &2 
a-2 &i 


(7) 


That is, two acids upon mixing do not influence each 
other’s dissociation when their concentrations are very 
nearly inversely proportional to their dissociation con- 
stants, or the ratio of their concentrations is equal to the 
reciprocal ratio of their dissociation constants. Let us 
suppose we wish to prepare i liter of a mixture containing 
xVmole acetic acid (£ 1 = i.8Xio~ 5 ) and mole glycollic 
acid (k 2 ^i$Xio~ 5 ). This can be done, without in- 
fluencing the dissociation, by combining solutions of 
acetic and glycollic acids having their concentrations in 
the ratio x\ :x 2 =i$ 11.8=8-3 :i. Hence we must mix 
8.3 volumes of glycollic acid with one of acetic acid of 
isohydric concentration (giving 9.3 volumes), and to 
fulfill at the same time the above conditions of concentra- 


8 3 

tion ~ liter of glycollic acid must contain X V mole, i.e., 


0.3 1 

be fr^Xo.i =0.112 normal, and — liter of acetic acid 
8.3 9-3 

must contain yV mole, that is, be 0.93 normal* 



66 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


The knowledge of isohydric concentrations is important 
for the reason that from the conductivities ki and k 2 , 
and the mixing volumes V\ and V 2 , the conductivity k of 
the mixture may be very simply calculated thus (by the 
rule of three): 

*i Fi + k 2 V 2 

* = fi+f 2 ■ 

Or, as Arrhenius expresses this: when two acids are 
present in a common solution, the conductivity may be 
calculated by introducing into the calculation for each 
resulting conductivity the concentrations based on the 
assumption that the acids distribute themselves in the 
aqueous solvent in the inverse ratio to their dissociation 
constants. The agreement of this statement with ex- 
perimental results has been proved by Wakeman 1 as well 
as by Arrhenius. And it deserves to be mentioned that 

r * r • j i - A 2 c 

in case of a mixture of acids the expression j ^ 

derived from the conductivity is not constant, as in the 
case of pure acids, but varies with c, so that the study 
of the conductivity becomes a valuable criterion of the 
purity of such electrolytes. 

In order to determine how strongly the presence of a 
second acid diminishes the degree of dissociation a x , of 
the first acid below the value /?i of the pure solution of the 
same concentration, we combine our two equations (6) 
(p. 62) into the expression: 

<xi 

ki (i-oq) 
k 2 ol 2 • 

(l-«2) 

1 Z^it§chr. physik. Chem., 15 , 159 (1894). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 67 


in which for weak electrolytes it is permissible to write 
1 ~ai = i —ct 2= 1, so that we may arrive at the convenient 
but nevertheless good approximation formulae: 


^2 

ai = T--a2 or 

R'2 R\ 


- ( 8 ) 


Now if w r e are interested in the acid 1, we employ the 
equations for the law of mass action, both for the pure 
and the mixed acid 1 : 

t «i , , N «i 2 / f k 2 

kl= ~ o~ == Z. ~~( a l c l +<^2^2) — ( Cl ~r J-C2 

1 pi i ~&i i— 

^ fa C 2 \ 

i—ai\ ki Ci) * 


Substituting again as above 1 — a'i = i — =1, we 

obtain: 


/?i 2 £ a ==tfi 2 -£i 


/ , k 2 Cq\ 

V + *1 Cl)’ 


or 


Pi . C 2 

= vjl + 7 . 

OL\ ^ ki Ci 


( 9 ) 


, v This then is the ratio in which the degree of dissociation 
of the acid 1, at the concentration c Vr is depressed by the 
addition of c 2 moles per liter of an acid having the dis- 
sociation constant k 2 . For the latter acid the analogous 
expression holds with interchanged indices. 

In order to form an idea of the magnitude of the values 
involved, let us consider the case of a mixture of 1 mole 



68 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


of acetic acid (£1 = 1.8X10 5 ) and 1 mole of cyanacetic 
acid (£2 = 37 oXio~ 5 ) per liter. In pure acetic acid we 


should have jSi=Vi.8Xio~ 5 = 0.00425; since £- = 205 
and — =1, we get 


/?i : = \/i + 2 05 = r 4.4 ; 


that is, the degree of dissociation is reduced to part, 

or from 0.00425 to 0.0003, or from 0.4% to 0.03%. In 
general we can see from the equation that the reduction 
of fti is greater the stronger the acid (£ 2 ) and the higher 
its concentration (c 2 ) is. 

To learn the counteraction of the acetic acid on the 
dissociation of the cyanacetic acid, we make use of the 
analogous expression 


^2 -«2 = 


44 --- 

> £2 C2 


.m'the v 

>tudy 


Here w r e see the -factor under the radical becomes e 

to 1+7^=1.0049 (instead of, as before, 206!), or 

r 2 . f a 

going back of the dissociation — is only about 0.25^ 


(instead of 1440% as before!); the general comparison 
is given by the equation: 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES 69 

The general result may be summed up in the statement 
that in a mixture of acids the acids mutually reduce their 
degree of dissociation, but the weaker acid is influenced 
very much more strongly. What has been said here about 
acids is, by analogy, absolutely true of all other weak 
electrolytes, in particular of bases. 

From the above it also follows that the farther two 
electrolytic solutions are removed from being isohydric 
the more also the conductivity of their mixture must be 
reduced as compared w T ith that of the unmixed solutions; 
for the minimum o of the mutual influence corresponds to 
the isohydric state. So having a solution a of an electro- 
lyte, it is possible to determine the isohydric condition 
of another solution b having ions in common with it, by 
adding the solution a to different concentrations of b. 
That concentration of b is isohydric with the given solution 
a, which by the addition of the latter brings about the 
maximum increase in conductivity. This deduction 
may under circumstances be useful for determining 
the degree of dissociation of a, when that of b is known 
and a is not directly determinable . 1 
. In a precisely similar manner the mass-action formulas 
for any more complex mixtures of electrolytes may be 
derived, as Arrhenius has made clear in his studies of the 
equilibrium relations between electrolytes . 2 Let the dis- 
cussion of a simple special case suffice here, a case of ex- 
treme importance to analytical chemistry, namely, that of 
two electrolytes, one very strong, the other very weak, 


1 Compare W. Bonsdorff, Ber. d. deutsch. chem. Ges., 3 G, 2322 (1903) ; 
Zeitschr. anorg. Chem., 41 , 132 (1904). 

2 Zeitschr. physik. Chem., 5 , 1 (1890). 



THE THEORY OF ELECTROLYTIC DISSOCIATION. 


so that by their mixture the dissociation of the strong 
electrolyte is only slightly influenced, while that of the 
weak is affected all the more. 

This case is realized in general whenever we mix a 
weak acid or a weak base with one of its neutral salts 
(strongly dissociated according to the rule, p. 26) or with 
a strong acid or base respectively. Then by equation 
(6) (p. 62), designating the weak electrolyte by 1 and the 
strong by 2, 


£1 = 


ai-Ci’(ai-Ci+a2’C2) 
(1 -ai)-ci 


i-«i 


(a 1 c 1 +a 2 c 2 ), 


the degree of dissociation <*i is greatly reduced according 
to the measure of the concentration c 2 . 

In order to obtain an approximation for this reduction, 
let us, as before, but with even greater exactness, write 
1 — #1 = 1, for the degree of dissociation of our weak 
electrolyte 1 in unmixed solution. This degree of dis- 
sociation, small by assumption, is made even smaller by 
the mixing. Further, for the sake of simplicity we shall 
assume that 0:2=1, or that the dissociation of the strong 
electrolyte is practically complete. Then we may also 
write our equation: 

h =<*1 +£ 2 ) =oti 2 Ci +aic 2 . 


Again, if the concentrations ci and c 2 are of the same 
order of magnitude, i.e., the concentration of the strong 
electrolyte is not very much smaller than that of the weak, 
we may in the latter expression drop the first summation 
as compared with the second . without introducing any 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 71 

great error, because a x , the degree of dissociation of the 
weak electrolyte in the mixture, is a small magnitude, 
hence its second power represents a magnitude of the 
second order. It follows then that approximately 

r • h 

Ki~o^i m C2, i.e., ct r i = — , 

C2 

*and the concentration etjCi of the non-mutual ion is 


aici=h— , (n) 

C 2 

or in words: the degree of dissociation a x of the weak 
electrolyte in the common solution of both electrolytes 
is directly proportional to the concentration of the strong 
electrolyte and inversely proportional to its dissociation 
constant. In order to point out the significance of these 
relations by way of several examples, let us consider the 
cases mentioned above of equivalent mixtures of acetic acid 
and alkali-acetate or of ammonia and ammonium salts; 
then for a x w r e must introduce the degree of dissociation 
of acetic acid and ammonia respectively, for the constant 
ki the value 1.8X10 -5 and 2.3 Xio -5 respectively. We 
find thus that in pure 1 -normal solutions of acetic acid 
and ammonia the degree of dissociation is respectively 
0.4% and 0.5%, while the same upon addition of 1- 
normal acetate and ammonium salt is depressed to 
0.0018% and 0.0023% respectively. These degrees of 
dissociation in the mixtures express at the same time, 
in the case of the acetic acid, the concentration of the H* 
ions, in the case of the ammonia, that of the OH' ions, 



72 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


and give us a measure of how very greatly the acid and 
basic properties, based on the concentration of these ions, 
are diminished. In the case of ammonia this fact was 
made use of long before its theoretical explanation was 
known, namely, in reducing by the addition of ammonium 
salts the power of ammonia to precipitate magnesium 
ions as magnesium hydroxide, or, practically speaking, 
counteracting it altogether. Likewise the reduction of 
the concentration of the sulphur ions in hydrogen sulphide, 
by increasing the concentration of H' ions through the 
addition of strong acids, is made use of in analysis to 
counteract the power of H2S to precipitate zinc. The 
equilibria appearing in connection with precipitating 
reactions will be further discussed later on. This driving 
back of dissociation, Arrhenius also proved experimentally 
for formic acid and acetic acid. As previously alluded 
to (p. 5), the inversion of cane-sugar is catalytically 
accelerated by acids, in proportion to the H* ion con- 
centration of these acids, which is shown by a compari- 
son of the catalytic action of varying acid concentrations. 
The following table gives the value of this catalyzing 
constant for several such acids and the influence upon it 
of additions of neutral salts. 

p indicates the reaction-velocity constant, i.e., the 
quantity of sugar inverted per minute, when the sugar 
possesses during the minute the concentration 1. The 
measurements were carried out at 54 0 , at which tem- 
perature k (acetic acid) =1-615 Xio" 5 an d ^ (formic acid) 
= i.83Xio - 4 , for calculating the H* concentration. 

For ¥ V norma l HC 1 it was found that <o=4.69Xio~ 3 ; 
this velocity of inversion is therefore brought about by 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 73 

an H‘ concentration = ^=0 0125 normal, since HC 1 in 
-gVN solution practically may be considered as completely 
ionized. 

A 0.25-N acetic acid has, since £ = 1.61 5Xio -5 , 
the H* concentration (H*) 2 = &Xo.25 or (H’) = 
v / o.25Xi.6i5Xio“ 5 = o.oo 2 normal, hence we should 
0.002 

have /)=4.6 oXio~ 3 — ! — — = o.7aXio“ 3 ; the value ob- 
r ^ y 00125 9 

served was p=o.j$Xio~ 3 : likewise for 0.25-N formic 

acid we calculate 0.25 X 1.83X1 o" 4 == 0.00678 

o 00678 

and £=4.69X10 =2. 54Xio“ 3 , while 2.55X10 3 

was the value found. In exactly the same manner the 
(H‘) values, and from these the t o calc . values, have been 
derived for the following mixtures, and the good agreement 
of Pcaic. and ^obs. proves that the basis of the (H‘) calcu- 
lation in the above equation agrees with the facts. 


0.25-N Acetic Acid 4 - c-Normal Sodium Acetate. 
(Arrhenius, 1889.) 


c—o 

0.0125 

i 

0.025 ! 

i 

0.05 

0.125 

0.25 

IO 3 j 0 obs. =0.75 
10 3 />calc. =0.74 

0.122 

O.I29 

0 0 

0 0 
d o' 

0.040 

0.038 

0.019 

0.017 

0.0105 

0.0100 


0.25-N Formic Acid + c-Normae Sodium Formate. 


c— 0 

0.025 

0.1 

0.25 

ioVoba. =2.55 

0.72 

0.24 

0.118 

IO^calc. =2.54 

o -75 

0.24 

0.117 


Another special case of electrolytic equilibria is that 



74 THE THEORY OF ELECTROLYTIC DISSOCIATION 


of a mixture of two electrolytes with common ions, having 
equal strengths (k 1 = k 2 ), as seems to be approximately 
the case with analogous salts. Then by equation (6) 
(P- 62) 


k\ = k 2 — 


(aiCi -ha2C 2 )oLiCi 
(1 —ai)ci 


(a'iCi+a2Co)o , 2 C2 
(1 —a2)c2 9 


that is, 


a\ a.2 k 

1 — a'x 1 — (x.2 (XiCi+a.2C2 9 


from which it follows ai=a 2 ~a and = — , — , or 

1 —a Ci ~ f c 2 

in words: the degree of dissociation in mixtures of 
electrolytes of like strength is equal to and of the same 
value as that which would correspond to each alone for 
the concentration (ci+c 2 ). 

From the theory of the electrolytic equilibria relations 
we can derive the explanation of a whole series of well- 
known manifestations. Theoretically in every case in 
which two electrolytes, that is their four ions, are present 
in the same solution, there must be formed by the inter- 
action of these ions some certain quantities of the four 
possible undissociated substances. So, for example, upon 
mixing KC1 and NaBr, there are present in the solution, in 
addition to these two undissociated salts, certain amounts 
of NaCl and KBr, resulting from the reciprocal interaction 
of their ions. Since, however, the tendency of all these 
four substances to dissociate is great, there cannot be 
formed any appreciable quantities of the new undisso- 
ciated substances. It is a very different situation if it 
so happens that one of the four possible substances 
possesses a very slight tendency to dissociate, or the 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 75 


reverse as we had probably better say here, in case ex- 
ceptionally strong atomic affinities are in action between 
any two of the ions participating in the equilibrium. 
This is, for instance, the case when we bring together the 
strongly dissociated substances HC 1 and Na-acetate. 
The possibility is then at hand that NaCl and H-acetate 
will be formed simultaneously; however, while NaCl is 
strongly dissociated, there exists a marked tendency to 
combine between the ions H* and acetate', which causes 
almost all of these ions to unite to form undissociated 
acetic acid, and results in a disappearance from solution 
of the particular ions as such. As one sees, this view 
contains the theory of the general experience that strong 
acids “ liberate ” the weak acid from the salts of weak 
acids, and likewise of course strong bases liberate the 
weak base from the salts of weak bases, or, in the language 
of the dissociation theory, transform their ions into the 
undissociated state. Wherever, then, an H' ion meets 
the anion of a weak acid, or an OH' ion the cathion of 
a weak base, the opportunity is made use of for both 
these ions to pass from the ionic state into undissociated 
substances. When we, for example, “set free” ammonia 
(£=2.3 Xio -5 ) from the strongly dissociated ammonium 
salt by the addition of strongly dissociated alkali-hydrox- 
ide, or carbonic acid (k = ^.o4Xio~ 7 ) from strongly 
dissociated sodium carbonate by strongly dissociated 
HC 1 , or hydrocyanic acid (k = i:^Xio~ 9 ) from strongly 
dissociated potassium cyanide by strongly dissociated acid, 
etc., we are doing nothing else than giving the ions of these 
weakly dissociated electrolytes the opportunity to unite 
and form uhdissociated electrolytes. Writing the equation 



7 6 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


for any such reaction, for example: 

K' + CN' -f- H* + CY = K“ + Cl' +HCN, 

we see that, similar to the case of neutralization, the 
unchanged ions remaining on both sides may be omitted, 
and the formula for the reaction becomes : 

H* + CN' = HCN. 

We have here evinced, then, a very striking analogy to 
neutralization, or, as we must express it in the sense of 
the ionic theory, the analog}' of the formation of undis- 
sociated water from its ions to the formation of a weakly 
dissociated electrolyte (HCN, H 2 C0 3 , NH 4 OH, H 2 S, etc., 
etc.) from its ions, or, put in the old way, to the liberation 
of weak acids or bases from their salts. 

The old formulation, still employing the last example, 
that potassium chloride is produced from potassium 
cyanide by the action of hydrochloric acid, is, strictly 
speaking, a distorted mode of expression, inasmuch as 
the constituents of potassium chloride continue in the 
same condition (ionic) after the reaction as before. The 
really essential part of the change is the formation of the 
undissociated weak electrolytes, precisely as it is the 
formation of water in neutralization. 


HYDROLYSIS. 

In the cases of aqueous solutions discussed thus far, we 
have left out of consideration altogether that there is. 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 


77 


peculiar to the solvent, water, as we saw above, a very 
small but nevertheless measurable and exactly known 
dissociation into the ions H* and OH 7 . We have, then, 
still to discuss how far and in what cases we are justified 
in assuming a participation of the water in the equilibrium 
between the electrolytes, and what may occur under 
those circumstances. For instance, in the simplest case, 
the solution of a strong salt such as NaCl, there is the 
possibility of the Na* ions combining with the OH' ions 
of the water to form undissociated NaOH, and the 
Cl' ions with the H* ions of the water to form undissociated 
HC1. We know, however, from the conductivities, 
freezing-points, etc., that neither of these two new com- 
pounds possesses to any sensible degree the tendency to 
assume the undissociated state, but that, on the contrary, 
they split up to a far-reaching extent into their ions. In 
consequence of this, neither of the ions of water is to any 
appreciable extent taken into custody by the ions of such 
a salt. The matter takes on a different aspect when, 
for example, the strongly dissociated salt of a very weak 
acid or base is involved; in such cases the ion of the 
weak acid or base present in the salt in large concentration 
finds an opportunity to unite with the ion of water neces- 
sary to form the weak acid or base. What is more, we 
are in position to determine, on the basis of the law of 
mass action, the extent to which this can take place. 
Taking under consideration a i -normal solution of potas- 
sium cyanide in water, we must have in this, in addition 
to the equilibrium between K* and CN' ions and undis- 
sociated KCN, the following three equilibria; 



73 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


1. K* ions and OH' ions (of water) with undis- 

sociated KOH; 

2. CN' ions and H* ions (of water) with undis- 

sociated HCN; 

3. H* ions (of water and the HCN formed) and 

OH' ions (of water and the KOH formed) 

with undissociated water. 

The K",OH' equilibrium, as well as the K‘, CN', cor- 
responds to strongly dissociated substances, while for 
the H‘, CN' equilibrium we have the equation (see pp. 63 
and 55) 


_ (H-)-(CN') 
(HCN) 


1.3X10- 9 , 


and for the H*, OH' equilibrium 


k w = (H*) • (OH') = 1.2 Xio~ 14 . 


Suppose this action of the water, the so-called hydrolysis, 
the result of which consists in the splitting up of part of 
a neutral salt into acid and base according to the equation 

KCN -f H 2 0 = KOH + HCN, 

has taken place to the extent x, so that x represents the 
fraction of each mole of neutral salt which has been split 
up in this way into base and acid. If c stands for the 
total concentration employed, then, according to the 
statements made, there would be present in the “ hydro- 
lytic ” equilibrium 

ivcKOH -h^cHCN 4- (1 -#)cKCN, 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 79 

Applying to these quantities the equilibrium equations 
of hydrocyanic acid and of water, 

i- We may practically identify the concentration of 
the undissociated HCN with the total concentration of 
the same equal to x-c, for hydrocyanic acid is an exceed- 
ingly weakly dissociated acid, whose degree of dissociation 
in addition is reduced by the presence of the many cyano- 
gen ions of the strongly dissociated KCN. 

2. The concentration of the CN' ions we may without 
marked error place equal to the concentration (i — x)c of 
the undecomposed KCN, since, for the reasons named, 
the cyanogen ions arising from HCN must be exceedingly 
few. 

3. The H* ion concentration, which is here required 
for the H‘,CN' equilibrium, we obtain with the aid of the 
water equilibrium, since we know the concentration of 
the OH' ions. This is practically equal to the total 
concentration of the (strongly dissociated) KOH, that 
is =x-c. Now since 

(H*) • (OH') = k w or = 

we can introduce all the values into the H*,CN' equilibrium 
equation and get 

(H-)-(CN') 1 — x 

k ‘~~ (HCN) “ x-c ~ kw ' 


K x 2 -c ' 

kg 1 — x 



v.- (**) 


In the form 



So THE THEORY OF ELECTROLYTIC DISSOCIATION. 


the analogy with the dilution law is very evident, and this 
hydrolytic dilution law can also be formulated in the 
words : 


(Cone. Acid) * (Cone. Base) 
(Cone, non-hydrolyzed Salt) 


= Hydrolytic Constant, 


just as we have 


(Cone. Cathions) ■ (Cone. Anions) 
(Cone, undissoc. Salt) 


= Dissociation Constant. 


Herewith we have an expression consisting entirely 
of known factors, namely, the known total concentration 
c and the constants of hydrocyanic acid (k 8 = i.^Xio~ 9 ) 
and of water (4 w = i. 2 Xio' 14 ), which enables us to 
calculate the degree of hydrolysis (x). 

Solving this equation for x, we get 



For the cases in which the ratio k u . : k s is small as 
compared with i, i.e., the constant of the w r eak acid is 
much greater than the water constant, this equation may 
be simplified without great error to the approximation 
equation : 

x= ^lvt (I3) 


The equation teaches us that the degree of hydrolysis is 
dependent on the ratio of the dissociation constants ol 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. Si 

the weak acid, whose salt we are considering, and that 
of the water. Hydrolysis increases, therefore, the weaker 
the weak acid or base contained in the salt is. In a 
qualitative way one can get a very good picture of the 
relations by the following considerations. 

The H* ions of the water act upon the weak anions of 
the salt with the formation of free undissociated acid; 
the place of the acid anions thus consumed is taken in 
equivalent amount by the OH' ions of the water that were 
formerly bound to the H' ions, producing a definite OH' 
ion titer of the solution. Thereby, however, the concen- 
tration of the H* ions (on account of the H* , OH' equi- 
librium, that must always be maintained) is reduced 
to such a small amount that no further free acid can be 
formed by their action. By the production, then, of 
the OH' ions or the free base the continuation of the 
hydrolysis is retarded (in that these ions suppress the 
hydrolytic action of the H' ions) and is finally brought 
to a standstill, this standstill setting in the later the less 
H' ions the acid anions require for the formation of 
undissociated acid, i.e., the weaker the acid is. 

The same is true of course, mutatis mutandis, for the 
salts of very weak bases, in which case the OH' ions of the 
water are the hydrolyzing and the H* ions the retarding 
ones. 

Starting with a salt hydrolyzed to the amount x, then, 
the concentrations (acid and base) are equal and equiv- 
alent, and the above equation holds, as Shields 1 demon- 
strated for KCN: 


1 geitscfrr. p&yslk- Chenfc, IS, 167 (1893),. 



82 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


Potassium Cyanide. 
(Shields, 1893.) 


c 

lOOX 

(l-*) 

0.947 

°-3 I % 

O.9 XlO“ 5 

0-235 

0.72 

I. 22 XIO -5 

0.095 

I . r 2 

i . 16 X 10— 5 

0.024 

2.34 

i 

1.3 X10- 5 
Mean: i.iXio~ 5 


The hydrolytic constant, which for immediate purposes 
we can derive on the basis of the law of mass action 


without any knowledge of the dissociation theory, never- 
theless represents the ratio k w : k 8 according to this theory, 
and therewith makes it possible to obtain the constant k 8 
of the weak acid, by means of (12), from the water con- 
stant ^ = 1.2 Xio -14 by dividing the same by the hydro- 
lytic constant. Accordingly, for HCN the same becomes 


1.2X10’ 14 
1 .1 Xio“ 5 


= 1.1 Xio 


-9 


while Walker found by direct 


measurement and in close agreement with it the value 
1.3X10- 9 . Similar to the significance given to the dis- 
sociation constant (p. 46), we may formulate the physical 
sense of the hydrolytic constant so that it signifies the 
half of that concentration at which the salt is just hydro- 
lytically decomposed one half, or (in case the degree of 
hydrolysis is small) the root of the hydrolytic constant of 
the concentrations of the products of hydrolysis in the 
i-N salt solution (when the salt contains only one very- 
weak ion). 

It is possible;' then, to obtain inversely . from the table 




EQUILIBRIA AMONG SEVERAL ELECTROLYTES &3 


(p. 55) the hydrolytic constant by dividing the water 
constant 1.2 Xio -14 by the dissociation constant given, 
that is, the degree of hydrolysis of the i-N salt solution 
according to (13) by dividing 1.1X10" 7 by the root of 
the dissociation constant. 


The constant 


(Acid) * (Base) 
(Salt) 


W alker 1 determined 


experimentally on mixtures of the very weak base urea 
with hydrochloric acid, by measuring the velocity of 
inversion p produced on the one hand by the pure acid 
(po) and on the other by the urea to which acid had been 

added. Then represents the fraction still having 

inverting action, and 1 tih- e P art th e acid bound as 

a salt of urea, as well as the urea itself thus bound, so that 

c ~(^ ~y S j £i yes f ree urea which, for the concentration 

c employed, is left after the formation of salt; conse- 
quently we should have 



Hydrolytic Constant. 


1 Zeitschr. physik. Chem., 4 , 319 (1889). 




&4 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


Normal Hydrochloric Acid + c-Normal Urea. 
(Calculated after Walker, 1889.) 


c 

9 

PO 

Cone, of 
Free HC1. 

1- — = 

<00 

Cone, of 
Salt 

Formed. 

«-.+*- 
9 0 

Cone, of 
Free Urea. 

Hydrol. 

Const. 

0 

0.0031 5 = ^3 

I 

— 

— 

— 

°-5 

0.00237 

o -753 

0-247 

0.253 

0.77* 

1 

0.00184 

0.585 

0-415 

0-585 

0.82 

2 

0.001 14 

0.36 

0.64 

I.36 

o-77 

3 

0 . 00082 

0.26 

0.74 

2.26 

0.80 

4 

0.0006 

0. 19 

0.81 

3-19 

o-75 

Mean: 0.78 


In a similar manner Walker’s measurements (re- 
calculated) give the following hydrolytic constants: 


Thiazole o . 00367 

Glycocoll o . 00425 

Asparagine 0.0079 

Thiohydantoin 0.0127 

Aspartic acid 0.0137 

Acetoxime o . 0196 

Acetamide 4.0 

Propionitrile 6.7 

Thiourea 10.5 


from which the dissociation constants given on p. 55 
for these basic-acting substances were calculated. 

Hydrolysis is outwardly recognized by the fact that 
a salt of neutral composition reacts in aqueous solution 
either acid or alkaline and not neutral, i.e., that the forma- 
tion of H' or OH' ions from the water is induced. In 
this particular the reaction of hydrolysis may possibly 
be most clearly represented thus, that the particular salt 
by means of its one weak ion acquires the opposite ion of 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . $5 

the water for the formation of the undissociated compound 
and thereby disturbs the H* , OH' equilibrium. Since 
this equilibrium, in consequence of the presence of water, 
must be maintained, there appears in place of the water 
ion which disappeared a quantity of the other water ion, 
to be calculated according to the equation 

(H0 = ^g 7 j and (OH , ) = ^- respective!}'. 

As a result the salt of a weak base reacts acid, and that 
of a weak acid alkaline. 

From the fact of this reaction it follows, on the basis 
of the law* of mass action, that hydrolysis in the case of a 
weak base is reduced by the addition of H* ions, in the 
case of a weak acid by the addition of OH' ions, for both 
these kinds of ions are reaction products of hydrolysis, 
and the increasing of the concentration of reaction prod- 
ucts always acts against the progress of the reaction. 
We may also picture this to ourselves thus: start with 
the consideration of the H‘ , OH' equilibrium and look 
upon the addition of OH' ions (in the form of any strong 
base) as a forcing back of the H* ion concentration of 
the water. This works against the reaction of these H* 
ions with the weak anion of the electrolyte, because the 
quantity of the undissociated weak electrolyte produced 
by the hydrolysis is proportional to the concentration of 
the hydrogen ions which come into consideration for this 
equilibrium. The same is true also, mutatis mutandis , 
of the hydrolysis of a salt of a weak base upon the addition 
of H’ ions. These deductions from the law of mass 



86 THE THEORY OF ELECTROLYTIC DISSOCIATION 


action may likewise easily be confirmed at the hand of 
experience. If, for example, we add strong caustic 
alkali, i.e., increase in this way very considerably the 
concentration of the OH' ions in a solution of potassium 
cyanide or ammonium sulphide, both of which reveal 
their hydrolysis not only by their alkaline reaction but 
also by the odor of undissociated hydrocyanic acid and 
hydrogen sulphide respectively, then the forcing back of 
the hydrolysis manifests itself in the disappearance of 
the odor of hydrocyanic acid and hydrogen sulphide 
respectively. Again, if we add strong acid to a solution 
of ferric chloride or iron alum, which react acid in conse- 
quence of hydrolysis and at the same time show the brown 
color of the undissociated ferric hydroxide (in colloidal 
solution), the brown color of the undissociated ferric 
hydroxide disappears more and more, giving place to the 
colorless condition belonging to the ferric ions. Gen- 
erally speaking, by forcing back hydrolysis by means of 
OH' ions or H‘ ions, those properties of a hydrolyzed 
solution that belong to the undissociated component 
disappear. 

As the equilibrium equation of hydrolysis shows, the 
hydrolytic decomposition can not only be forced back- 
ward by the addition of either H' or OH', but also by 
the addition of the undissociated product of hydrolysis. 
Thus, for instance, the acid reaction of a solution of 
aniline hydrochloride is destroyed by an excess of aniline, 
and Bredig 1 was enabled by this device to measure the 
ionic mobilities of salts subject to hydrolytic decomposi- 
tion. 


‘ 1 Zeltschr. physik. Chem., 13 , 214 (1894). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. S; 

In reality, according to the theory, all sails should be 
subject to hydrolytic decomposition, and the equation 
(13) (see p. 80) for hydrolysis enables us to determine 
quantitatively the degree of decomposition as soon as 
we are in possession of the dissociation constant of the 
weak constituent of the salt. It turns out that the degree 
of hydrolysis of weak electrolytes, whose dissociation 
constant is of the order of magnitude of that of acetic 
acid, is still exceedingly slight, so that a 0.1 -normal 
solution of sodium acetate is hydrolyzed only 0.008%, 1 
as shown by the investigation of Shields. 2 He measured 
a reaction velocity which is proportional to the concen- 
tration of the OH' ions, namely, the saponification of 
ethyl acetate, whereby he determined the OH' ion con- 
centration and therewith the hydrolysis of salts of weak 
acids. His results are contained in the table, p. 91, as 
well as Walker’s 3 calculated degrees of hydrolysis, which 
he obtained by means of the dissociation constant of the 
weak acid, this constant having been determined by 
conductivity measurements. 

As can be seen, noteworthy degrees of hydrolysis are 
to be expected only in the case of salts of extremely w r eak 
electrolytes, such as those enumerated in the table on p. 
55. All the salts of the acids and bases mentioned there 


1 According to equation ( 13) : 

a/JL IO %W o~^6 7 X io-» = o.8 X 10-*. 

y 10 i.sxio - 5 

2 Zeitschr. physik. Chem., 12 , 167 (1S93). 

*Ibid., 32 , 1 37 (1900), and Joum. Chem. Soc., 77 , 3 (1900). 



88 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

are hydrolyzed to such an extent that one can demonstrate 
by indicator reactions the presence of measurable quan- 
tities of OH' or H* ions. One frequently meets with 
the view that the determination of this OH' or H* 
concentration, or, what is the same thing, the alkali 
or acid titer, is possible by means of alkalimetric or 
acidimetric titrations respectively. But this is impossible, 
for the reason that the. ionic reactions take place with 
an immeasurable velocity, and in consequence, as the 
H* or OH' titer in the hydrolyzed solution changes, new 
equilibria of the various ionic concentrations establish 
themselves at once. An illustration, at the same time of 
great importance to the chemistry of our daily life, will 
help to make the case clear. Consider a solution of 
borax, which on account of the weakness of boric acid 
(see p. 55) is hydrolyzed and hence gives an alkaline 
reaction; in this for the moment we may attempt to 
convert these hydroxyl ions into water by the addition 
of hydrogen ions. This would at once raise the H* ion 
concentration to a greater value than is in keeping with 
the equilibrium H' , H2BO3' of boric acid. The excess 
of H* ions would then be taken up again by the borate 
ions forming undissociated boric acid, whereupon a 
fresh quantity of hydroxyl ions would again have to be 
formed, owing to the H* , OH' equilibrium, or, in other 
words, the alkaline reaction, that we attempted to destroy 
by the addition of acid, persists in spite of it. In fact, 
this continues as long as appreciable quantities of borate 
ions are present in the solution, sufficient to take pos- 
session of the H' ions added with the acid. On this 
behavior is based the possibility of choosing borax for 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 89 


neutralization, as is not infrequently done in alkalimetry, 
instead of alkali-hydroxides. Likewise potassium cyanide 
might be used, but for other reasons it is not feasible. 

The application of soap, soda, and borax in the house- 
hold is also essentially based on the fact that these Xa 
salts of very weak acids, as a result of hydrolysis, give 


solutions of appreciable but yet of^ sj^h^sma^QEL', con- 
centration that l tfiffi£^|Bti^®t|(iai fisl |^il ; innjDt5hg, 
though still stimcrent ro make their swelling and tat- 
solvent action effective. 


mm 



i is. ^ 


d, without stliei-efore 


must be still more strdf 
the concentration of the H' cr OH' ions becoming greater 


than in the previously discussed case. The mass-action 
equation for the hydrolysis of such a salt can likewise be 
very simply derived. Take a salt such as ammonium 
cyanide or aniline acetate of the concentration c , and 
indicate, as before, the degree of hydrolysis by x, that is, 
the quantity of free acid and base formed by the water 
per mole of salt, then in the hydrolytic equilibrium we 
have: 


x-c free base free acid + (i —x)c non-hydrolyzed salt. 


Indicating further the dissociation constant of the 
base by k b , of the acid by k 8 , and the ionic product of 
water by k w (at ordinary temperature io~ 14 ), then the 
following conditions of equilibrium must be fulfilled at 
the same time: 


i. 


For the base: 


(Cathions) * (OH') _ 
(Undissoc. Base) bl 



t)0 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


n 


For the acid: 


(Anions) • (H‘) 
(Undissoc. Acid) 



3. For the water: (H*) -(OK r ) = k w . 


On the strength of similar considerations as above 
(p. 79), we may place the concentrations of the undisso- 
ciated base as well as the undissociated acid equal to their 
total concentration (x-c), and further, as experience has 
shown (of the four electrolytes here present only the 
neutral salt is to be considered as strongly dissociated), 
we may with useful approximation assume the concen- 
tration of the basic cathions as well as of the acid anions 
as equal to the concentration of the salt (1— x). We 
obtain then 


(OH')-(i -afl-c : 
x-c 

(H')-(i —x) c 


and by multiplying these two values, the relation 

or (compare pp. 79, 80) 

k w (Cone. Acid) • (Cone. Base) 
k^kT (Cone. Salt) 2 ’* * (l5) 

after having introduced for the product (H*)* (OH') its 
value k w . We get for the hydrolysis of such a salt, consisting 
of two weak components, the interesting result that the 
hydrolyzed portion is entirely independent of the concen- 
tration of the salt, since this concentration falls out of the 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 91 


final equation. The correctness of this relation has 
been experimentally confirmed by Walker at the instiga- 
tion of Arrhenius. The accompanying table on hydroly- 
sis contains these results taken from the fundamental 
research of Arrhenius . 1 


Degrees of Hydrolysis. 
(Shields, 1893, and Walker, 1900.) 
Salts of Weak Acids. 



Shields, Found. 

j _ _ . ' 

' Walker, Calc. 

0 t-ML Sodium acetate 


0 . coS% 
0.06 

0.1-N. Sodium bicarbonate 


0 i-N. Sodium hvdrosulphide 


0.14 

0.84 

°3 

0.96 

0 i-N. Sodium metaborate 

; 

0.1-N. Borax =2NaB0od-B o 0 3 

0.1-N. Sodium cyanide \ 

; about 0.5 

0 . i-N. Potassium cyanide J 

0.1-N. Sodium phenolate ) 

1 . 1 

! ; 

0. i-N. Potassium phenolate J 

3 *° 

3 -o 


Salts of Weak Bases. 

Vgo— N. Aniline chlorhydrate 2.6% 

V32— N- ^Tohiidine chlorhydrate. .. . 1.5 

1 / 32 -N. 0-Toluidine chlorhydrate. ... 3.1 

1 / 33 -N. Chlorhydrate of urea 95-° 


Aniline Chlorhydrate. 
(Bredig, 1894.) 


- T 

C 

100# 



t i — x ^Aniline 
c x 2 ~~ k w 

3 2 

2.63 

45X10 3 

64 

3-9° 

40X10 3 

128 

5-47 

40X10 3 

256 

7.68 

40X10 3 

512 

10.4 

42X10 3 

1024 

14.4 

42X10 3 


1 Zeitschr. physik. Chem. 7 5 , 18 (1890). 




92 THE THEORY OF ELECTROLYTIC DISSOCIATION . 


Aniline Acetate. 


(Arrhenius and Walker, 1890.) 



100X 

12.5 

54-6 

i 25 

55-3 

5 ° 

56*4 

100 

55 - 1 

200 

55-6 

400 

55-4 

Soo 

5^-9 

Mean: 55. 7 1 


Of inorganic salts those of the following weak cathions 
show hydrolysis with acid reaction 2 when associated 
with strong anions: 

■ Be'*, Hg 2 ” Hg“, Cir*, Al**‘, Cr***, Fe***, Mn***, Sn’*, 
Sn****, Sb-, Bi*“, U:::. 

With weak anions we usually have great insolubility, 
the formation of basic salts, or inappreciable dissociation 
(Hg**). Alkaline reaction is shown by the weak anions, 

H 2 B0 3 ', P0 4 '", HPO4", HS', S", CO3", ’HC0 3 ', 
Cr0 4 ", Si0 3 ", SO/', CIO', N0 2 ', 

w r hen associated with strong (alkali) cathions; with weak 
cathions we have in all cases insolubility. 



k w 1.2 x 10 14 

H'ka 4.9 X IO —10 Xl.8X io~ 5 I ' 4 ' 


Compare equation (15), p- 90. 

2 Concerning the quantitative relations, compare the thorough and 
comprehensive study of H. Ley, Zeitschr. physik. Chem., 30 , 193 (1899). 




EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 93 


The other question, whether such a salt reacts alkaline 
or acid, may according to the above mathematical re- 
lations for k b and k s be very easily decided. If we recall 
that a neutral reaction means the same as the presence of 
equivalent amounts of H* and OH' ions, or, in the form 
of equations, 

(H-) = (OH') or 

then by analogy the basic reaction must be formulated 
thus: 

JSl <t 

(OHO ’ 

and the acid reaction: 

(OHO ’ 

and so we may express the value 

(HQ 

(OH') 


according to equation (14) (p. 90) in terms of the ratio 

k 8 - 

for, as we see. 


h_ (HQ 
K (OHO 


(16) 


Hence we obtain acid reaction when, as with aniline 
acetate^ 

k s > k b/i 



94 THE THEORY OF ELECTROLYTIC DISSOCIATION. 
basic (alkaline) reaction when, as with (NH 4 ) 2 C 03 or 

(NH 4 ) 2 s, 

k& ^ kb j 

and neutral reaction when, as is approximately the 
case with (NH 4 ) 2 C 2 H 3 02 , 


k s k' 0 


The hydrolytic relations of a salt with two weak ions 
will also establish themselves if we add to a solution of a 
salt of one-sided weakness a salt that has a second w T eak 
ion of opposite nature. To cite an example, consider 
the hydrolysis of ammonium chloride, which is limited 
upon reaching a certain (in this case very small) H’ ion 
concentration. If we add another salt with a very w T eak 
anion, for example KHS or K 2 C0 3 , then its anion HS', 
or C0 3 ", will consume the H* ions, formed by the hy- 
drolysis of the ammonium chloride, for the production 
of the undissociated acid H 2 S or H 2 C0 3 respectively. 
Herewith the previously existing check, which interfered 
with the progress of the NH 4 C1 hydrolysis into NH 3 +HC1, 
is removed. The result is that the first NH 3 -producing 
hydrolysis goes on and the newly formed H' ions continue 
to be bound by the HS' or C0 3 " ions respectively, etc., 
so that the undissociated products of hydrolysis of a 
w T eak pair of ions must reach much higher concentrations. 
In the case selected this becomes very evident; for w hil e 
a solution of NH 4 C1 is so little hydrolyzed that there is no 
sign of the odor of ammonia, the odor is in evidence at 
once on the addition of K 2 C0 3 or NaHS solution. In 
the latter case the odor of H 2 S also appears as an indica- 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 95 


tion that the HS' ions have executed their H '-binding 
action. 

Such action is also made use of in analysis, for instance 
in the precipitation of Al‘*‘ and Fe'“ by means of an 
acetate (see.p. 142). 

The hydrolytic relations may in addition be viewed 
from another side, by considering the possibility of 
arriving at and examining the equilibrium of a reaction 
not only by starting with the reacting substances (salt + 
water), but also by starting with the products of the 
reaction (acid+base). 

We arrive at the equilibrium in the latter way by adding, 
we shall assume gradually, an equivalent amount of 
strong base to, for example, a weak acid (dissociation 
constant = k 8J ionic concentration in pure solution =\/&«, 
equivalent for anions and H’ ions). Thereby we disturb 
the equilibrium of the weak acid, for we consume the 
H' ions of the same, which must form undissociated 
water with the OH' ions of the added base. In conse- 
quence of which, in order to maintain the ionic product 
k S} a quantity of acid anions must be produced equivalent 
to the consumed H‘ ions, and these in turn again reduce 
more and more the H' ion concentration, on account of 
the dissociation equilibrium of the weak acid, until 
finally this has become as small as is demanded by the 
H*, OH' equilibrium. 


AVIDITY. 

The phenomenon of hydrolysis, discussed in the 
preceding pages, is fundamentally nothing more than 



9 6 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

an interesting special case of a general and important 
equilibrium problem. Long before the time of the 
dissociation theory, this problem had been investigated 
thermochemically by Thomsen 1 and volume-chemically 
by Ostwald, 2 as the distribution of a base between two 
acids and an acid between two bases, and was termed by 
the former the avidity of acids and bases. 

In order to dispose of the theory of the question, let 
us consider the case of having mixed in a liter of solution 

b moles of a strong base (BOH), 

ci moles of the acidx (HAi), and 

c 2 moles of the acid 2 (HA 2 ), 

and let b <Ci+c 2 , i.e., the amount of the base is insufficient 
to neutralize both acids. If now we call x the fraction 
of each mole of base which reacts with the acidi, then 
(i —x) is evidently the fraction remaining for the acid 2 . 

In the resulting equilibrium there will be formed 
and be remaining: 

Salt BAi : bx ; Acid HAi : Ci — bx; 

Salt BA 3 : 6(i -*); Acid HA 2 : c 2 -b( i -*). 

If both acids are weak, and both salts, as is usually 
the case, almost completely dissociated, then the 
ionic concentration (A i)=bx, (A 2 ') = b(i —x), and the 
remaining portions of the acids may, on account of the 
presence of their salts (see pp. 70, 71), be considered 

1 Thermochemische Untersuchungen. Leipzig, 1884. 

2 Journ- prakt, Chem. (2), IS, 328 (1878). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 97 

as essentially undissociated; their small hydrogen ion 
concentration (H*) in the solution has of course the same 
value. Accordingly, the dissociation constants ki and k 2 
of the acids are represented by the following expressions: 

(H-)(AiQ (HQ-to 

1 (HAi) d-bx’ 

, (HQ(A 2 ') (H-)ft(i-y) 

2 (HAa) c- 2 ~ b(i— x) 

so that the quotient 

ki (Salt'BAi) (AcidHA 2 ) x c 2 —b( i—x) 

k 2 ~ (Salt BA 2 ) (Acid HAi) ~i—x c-^—bx 

or \ 

(Salt BAi) k x (Free Acid HAQ | 

(Salt BA 2 ) k 2 (Free Acid HA 2 ) J 

represents the equation from which the distribution ratio 

X 

of the base - — — may be calculated for any case, as a 

function of the acid dissociation constants and their 
quantities (c 1 and c 2 ). The solution of this general 
equation leads to the expression 

(K—i)b -\-Kc\ d~c 2 

± \^[(K — i )6 -H Kci + c 2 f- 4 (K ~ i )Kbc 1 
* — ' < l8 > 



98 THE THEORY OF ELECTROLYTIC DISSOCIATION . 
and 


(K—i)b — Kci—C2 


i — #=- 

in which 


: v [(iv — i )b ~r Kc\ + C2] 2 — 4(K — i )Kbci 
2 (K-i)b ’ 


K= 


h 


The simplest case, investigated by Ostwald and theo- 
retically calculated by Arrhenius, was the one in which 
equivalent quantities of a salt and a second acid were 
mixed, or, since we may look upon the salt as made up of 
one equivalent each of base and acid, it follows that in 
our general equation we must write b=c 1 =c 2 . That 
means that the above quotient (17) becomes 


ki x b-x x 2 

k 2 1 — x b(i — x) ( 1 — x ) 2 



Since in the case of equal concentration the ionic 
concentrations of pure (unmixed) acids are to each other 
as the roots of their dissociation constants (p. 47), we may 
with Arrhenius also express this equation thus, that both 
acids divide themselves between the base in the same 
ratio as their degrees of dissociation would be, if each 
were present alone in the volume considered. Arrhenius, 
derived this for the case of a strong and a weak acid. 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 99 

In both cases Arrhenius compared his calculations with 
Ostwald’s measurements, and found an excellent agree- 
ment, as the following examples show. They refer to 
solutions which contain of each of the three substances 
0.33 moles per liter, resulting from the mixing of- 1 liter 
of each of the three normal solutions. The figures in 
the table give the value x, that is, the portion of the first 
and stronger acid consumed per mole of base. 


Distribution-ratio of Two Acids between One Base. 
(Observed by Ostwald, 187S; Calculated by Arrhenius, 1SS9.) 



Observed. 

Calculated- 

UNO, : CUCH.COOH 

0.76 

0.70 

HC 1 : Cl 2 CH.COOH 

0-74 

0.70 

CCI3COOH : CkCH.COOH 

0.71 

0.70 

CHC 1 XOOH : CH 3 CH(OH)COOH 

0.91 

o -95 

CCI3COOH : CHodiCOOH 

0.92 

0.92 

CCI3COOH : HCOOH 

0.97 

0.96 

HCOOH : CH 3 CH(OH)COOH 

°-54 

°-5 5 

HGOOH : CH3COOH 

0.76 

0.77 

HCOOH : Ca^COOH 

0.80 

o -79 

HCOOH : iso-CgHyCOOH . 

0.81 

0.79 

HCOOH : C,H 5 COOH 

°-79 

0.80 

HCOOH : CH>OHCOOH 

0.44? 

0.54 

CH3COOH : CgHyCOOH 

o -53 

0-53 

CH3COOH : iso-CjHyC O OH 

°'53 

0-53 


The more complicated case, the interaction of a weak 
acid with any concentration, not equivalent, .of the 

salt of another weak acid (b = ci ;> £2^ , was investigated 

by Wolf 1 by means of conductivity studies, and was 
found to agree splendidly with equation (18), easily 
simplified for this case. 

Zeitschr. physik. Chem., - 10 , 226 (1902). 



loo THE THEORY OF ELECTROLYTIC DISSOCIATION. 


INDICATORS. 

A number of weak electrolytes possess the peculiarity 
of having a very different color for the undissociated 
part and its ions, or of having only one of the two colored. 

If the substances possessing a different color in the 
undissociated and dissociated condition are quite weak 
electrolytes, they may be employed as “ indicators.” 
This prerequisite makes it impossible for anything but 
bases and acids to be classed here, salts being excluded, 
since as a rule they are strongly dissociated. 

In using indicators the purpose is to recognize whether 
a solution is neutral, or whether it contains H' or OH' 
ions in excess. If the indicator is (i) an acid, then 
H* ions work against its dissociation, that is, make its 
anions disappear, while OH' ions (by “ salt formation ” 
with the same) produce its anions; if on the other hand 
the indicator is (2) a base, then H* ions react with the 
same to form salt, that is, produce its cathions, while OH' 
ions force back its dissociation, or, in other words, trans- 
fer the cathions to the undissociated substance. 

It is scarcely necessary to add that, as far as the color 
change is concerned, only the anions of the acid or the 
cathions of the basic indicators come into question, since 
neither the H’ nor the OH' ion possesses color, as is 
proved by the existence of colorless acids and bases 
(see p. 13). 

The theory of the action of indicators may be developed 
by means of the recently derived equations, if we only 
remember that here also we have the competition of 
two acids for one base, of which the indicator acid is 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES, iol 


one, or the competition of two bases for one acid, of 
which the indicator base is one. 

An example will make the matter clear. Suppose we 
desire to titrate dichloracetic acid (ki =5.1 Xio" 2 ) by 
means of the indicator acid />-nitrophenol (i 2 = i.2Xio“ 7 ). 
(We shall see later that in this case it would be better 
to choose another indicator.) Nitrophenol is a weak 
acid, which by itself in an aqueous solution is so little 
dissociated that the intensely yellow color of the nitro- 
phenol anion is scarcely perceptible. Upon the addition 
of a base, i.e., OH' ions, it is, however, as good as com- 
pletely converted into (H 2 0 and) nitrophenol salt, which 
means nitrophenol anions, and so gives rise to yellow 
color. 


Let us assume that we are titrating 100 c.c. of a dichlor- 
acetic acid solution of the concentration 2 c x (per liter), 
to which a quantity of nitrophenol has been added such 
as to give the latter the concentration 2 C 2 (per liter), with 
an equally strong base, say potassium hydroxide, likewise 
of the concentration 2C1. We now add an amount of 
alkali which is exactly equivalent 1 to the dichloracetic 
acid, and apply the general avidity equation (see p. 97, 
equation [18]), in order to determine how much of the 
alkali goes to the dichloracetic acid (acidi) and how 
much to the nitrophenol (acid 2 ), since these are the two 
acids competing for the base. 


ki 5.1 Xio 2 . 

For the above case A = y- = rr — 31=4.25 Xio~, and 

ko 1. 2X10 7 D ’ 


also b=c x . By introducing these values into equation 


1 Thereby the original concentrations 2C1 and 2 c z sink to half the 
values, Ci and c 2J on account of the doubling of the volume. 



102 THE THEORY OF ELECTROLYTIC DISSOCIATION. 
fiSj and neglecting i as compared with K, then on the 

c i 

assumption that the indicator concentration = 
which will be about the equivalent of the quantities prac- 
tically employed, we find in round numbers for 

the portion i—x, which is the portion left for the nitro- 
phenol by the added base. That is, 999995 millionths 
of the dichloracetic acid is neutralized by the concentra- 
tion Ci of the base, and — — — falls to the concentration 
ioooooo 

c 2 of the nitrophenol, which we assumed to be=^^. 

Hence of the nitrophenol is first neutralized, i.e., 

0.005^2 is the titer of the yellow nitrophenol anions. If 
now one more drop ( = approx. 0.04 c.c.) of the titrating 
alkali of the concentration 2c x is added, there is brought 
into the solution 0.04 X 2C\ = 0.080^! millimole of base, 
while there was still left in the approximately 200 c.c. 
. _ . 200X<-Ci , 

of the titrating mixture = o.ooiCi millimole of 

IOOOOOO 

free dichloracetic acid. Even if straightway all the 
dichloracetic acid would now be neutralized, there would 
still be left 0.079^1 = 79^2 millimoles of alkali for the 
neutralization of nitrophenol, of which almost all of the 
2ooc 2 millimoles in the 200 cx. are still present unneu- 
tralized. There will be formed, then, in our 200 cx. of 
titrating liquid (outside of hydrolysis) about jgc 2 miHir 
moles of nitrophenol salt, which means a titer of 

7QC‘> 

— -= approx. 0.04c 2 normal as to yellow nitropheiiol 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 103 

anions. With one drop of alkali less the titer = 0.005^2; 
this drop then increased it about 80 times. 

The calculation shows that it is of importance to 
choose an indicator with a somewhat small dissociation 
constant £ 2 > so that the same does not take up appreciable 
quantities of the base before the stronger acid is neutral- 
ized; the wreaker the acid to be titrated is, the much more 
w r eak the indicator must be, since by the avidity equation 
the ratio of the constants is the determining factor. 
Likewise the indicator concentration must not be too 
great, so that the mass action does not compensate for 
the weakness; on the other hand, it is not advantageous 
to take the concentration too small, so that the first drop 
of excess of the titrating solution may be most effectively 
utilized to form the colored indicator ions. As a rule, 
then, it is best to take an amount of indicator equivalent 
to the quantity of acid or base contained in one drop of 
the titration liquid. This may be accomplished by 
adding to the reaction mixture one drop of an indicator 
solution made up equivalent to one of the titration liquids. 

The selection of an indicator having a very small 
dissociation constant, such as phenolphthalein, makes 
it necessary that the base used in neutralization be strong. 
Suppose, for example, we titrate our acid w T ith ammonia, 
using phenolphthalein as indicator, then in the first stages 
of the titration up nearly to the neutral point w r e wrould 
have formed essentially the ammonium salt of the acid 
to be neutralized, and so the first traces of excess of 
ammonia would find themselves in the presence of a 
large quantity of ammonium ions. These (according to 
p. 71) can bring only very few OH' ions into the solution 



104 THE THEORY OF ELECTRGL YTIC DISSOCIATION. 

and would not be able to dissociate the indicator acid 
to any marked extent, or, in other words, the ammonium 
salt of the indicator becomes far-reachingly hydrolyzed, 
which amounts to a non-formation of its colored ions. In 
such cases, therefore, it would require a considerable 
excess of ammonia to gradually form sufficient indicator 
ammonium salt to show the color change distinctly, that 
is, the turning-point would not be sharp. It is evident, 
then, that no very weak indicator acid can be employed 
in the titration of weak bases, but that we must employ 
relatively strong ones (always, however, much weaker 
than the acid taking part in the titration). Such rela- 
tively strong indicator acids are, for instance, nitrophenol 
and methyl orange. On the other hand, a very weak acid 
(or a strongly basic) indicator such as phenolphthalein 
is necessary in titrating weak acids, so that the indicator 
in its competition with the other acid does not successfully 
interfere (i.e., form anions) before the neutralization 
of the latter. 

From this we may derive the following rule for the 
selection of the indicator acids (indicator bases are not 
in use). If we have to titrate: 

1. Strong acid and strong base — indicator at 

random; 

2. Strong acid and weak base — indicator strongest 

(methyl orange, nitrophenol); 

3. Weak acid and strong base — indicator weakest 

(phenolphthalein, litmus) ; 

4. Weak acid and weak base — to be avoided, 

since the color change with every indicator 

is not sharp. 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 105 


The importance of the exact knowledge of the electro- 
lytic equilibria for these relations is shown by the following 
question: Which indicator is to be chosen in order to 
titrate hydrochloric acid, containing ammonium chloride, 
with potassium hydroxide? 

Having here case 1, one might think it permissible 
to choose at random and employ phenolphthalein as 
indicator, since it gives the sharpest color change, and 
potassium hydroxide is ordinarily easily titrated with it. 
In this case, however, we must remember that the NH^ 
ions of the ammonium chloride addition destroy the 
large OH' concentration of the first excess of potassium 
hydroxide by forming undissociated ammonia, and that 
phenolphthalein in order to form salt (red coloration) 
requires much OH'. Therefore, in order to get a sharp 
end-point it is necessary to use a stronger acid indicator; 
hence case 1 must be modified thus: if the solution 
contains a salt with a weak cathion, then the equilibrium 
relations correspond to those of case 2; if it contains a 
salt with a weak anion, then they become those of case 3. 

HETEROGENEOUS ELECTROLYTIC EQUILIBRIA. 

It is scarcely possible to treat here exhaustively every 
phase of electrolytic equilibria; for that reason let us 
consider only one more case, one which is of fundamental 
significance to analytical chemistry, namely, that the 
concentration of an electrolyte is limited by its solubility. 
As to form such equilibria become simpler, because the 
concentration or active mass of the substance, which is 
present in the state of saturation and is kept in this con- 



IO 6 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

dition by the presence of the substance in the solid state, 
becomes constant. We have learned to recognize the 
electrolyte <£ water ” as a substance of never varying 
concentration, whose ionic product, the water constant 
is unchanging, because the undissociated water always 
has very nearly the same concentration. Precisely the 
same holds for the ionic product of any substance whose 
saturation is maintained, for then, according to Nemst’s 
partition law, the solution contains under all circumstances 
(at constant temperature) an invariable 1 concentration l 
of undissociated electrolyte. According to the law of 
mass action, the product of its ions M* n and A ,m is : 

(M*) n • (A') m = k 

in which, according to Ostwald, L stands for cc solubility 
product ” (or the ionic product). This equation show's 
us at once what the effect is of the addition of an electro- 
lyte having, for instance, M* in common: we have formed 
immediately more undissociated substance M n A m , but 
since the solution was previously saturated with the same, 
it precipitates until the ion A', not in common w r ith the 
addition, is reduced so far that its concentration, following 
the above equation, has become 



that is, upon increasing M*, A' becomes correspondingly 
smaller than originally. 

1 It is true the presence of other electrolytes changes somewhat the 
solvent medium. Compare Arrhenius, Zeitschr. physik. Cfiem., 31, 
T 97 (i899)- 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 107 


This formulation we owe to Xernst, 1 and a thorough 
experimental testing to Noyes, 2 as well as to Goodwin, 3 
who employed an entirely different method from that of 
Noyes, namely, the measurement of concentration chains. 
The quantitative agreement between theory and experi- 
ment still suffers from the uncertainty with which the 
ionic concentrations in strong electrolytes (seep. 12 1) are 
burdened, but nevertheless the same is sufficient to show 
the correctness of the law of solubility influence. The 
following small table taken from Noyes’s measurements 
teaches especially that the equivalent addition of the one 
or the other ion has the same action on the solubility of a 
binary electrolyte: 


Solubility of Thallous Chloride (25°). 
(Noyes, 1S9C.) 


Concentration 
of the 
Addition. 

Addition* 

TlNOj 

Addition: 

HC1 

0 

0.0161 

0.0283 

0.0560 

O.147 

0 . 0083 
0.0057 
0.0033 

0 . 00S4 

0.00565 

0.0032 


With very insoluble salts, such as AgCl, the quantity 
in solution may be looked upon as very nearly completely 
dissociated, on account of its great dilution, so that only 


1 Zeitschr. phvsik. Chem., 4 . 372 (18S9). 

2 Ibid-, 6, 241 (1S90). 

* Ibid.* 13 ,. 588 (1894k 




xoS THE THEORY OF ELECTROLYTIC DISSOCIATION. 

a minimum portion of the quantity dissolved is undis- 
sociated substance, while the major portion is present 
as ions. If, for example, we reduce to to the Ag* ion con- 
centration of a saturated AgCl solution by the addition 
of HC1 (i.e., CF), it is practically equivalent to reducing 
the total amount of silver (Ag* ions+undissoc. AgCl) 
present in the solution to to, even, if strictly speaking, the 
quantity of the undissociated AgCl is not changed at all 
in its concentration by the addition. 

Let us now consider conversely a mixture of two 
electrolytes, each of which contains an ion of a difficultly 
soluble substance, for example AgN0 3 and KC1, and 
discuss the conditions, when and how much of the diffi- 
cultly soluble substance is formed. In any case the 
quantity of undissociated AgCl becomes 


k • (AgCl) = (Ag*)(CF). 

If either the Ag* or CF ions or even both ions are very 
dilute, so that the concentration (AgCl) formed from 
these may be smaller than that required for saturation, 
then no precipitate of AgCl is produced, for not until we 
have (Ag*)(CF)>Z can solid AgCl separate; from there 
on, however, an increase of either the one or the other 
kind of ion no longer produces an increase of the ionic 
product, since the value L is its maximum value. If 
both ions are present in equal quantity, then 
fAg') = (CF)=\/L, the solubility of the silver chloride in 
water (more exactly, after subtracting the undissociated 
AgCl). The value of this solubility product L , which 
of course varies for different substances, is of fundamental 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 109 


significance in the formation and conversion of pre- 
cipitates. 1 Several examples will show this. 

If AgCl reacts with KI, Agl and KC1 are formed. 
How far does this reaction go? In the common solution 
of the four substances the Ag* , Cl' equilibrium demands 


(Ag-) = 


^AgCl 

(CIO ’ 


and the Ag' , V equilibrium, 


hence we have 


(Ag*)= : 


(10 5 


Z Ag q_(C10_ 

La s i (10 


That is, the interaction proceeds until the ratio of the 
(CIO • (10 ions has reached the constant value K, the 
quotient of the two solubility products. If therefore we 
use for the precipitation of any Ag' solution a potassium 
chloride and iodide mixture in the ratio (Cl') : (I') = 
i A g ci ; ^Agi, then the difficultly soluble salts also 
precipitate in this ratio until the remaining Ag' con- 
centration satisfies both solubility products. If we 
further add KI to this equilibrium, this would disturb the 
equilibrium ratio (Cl') : (I'), and consequently the reaction 
AgCl+I' = AgI + Cl' takes place, until the old value 
(Cl') :(I') is attained; but also the reverse, the addition of 
KC1 converts some Agl into AgCl with the production 


See in particular Findlay , Zeitsciir. physik. Chem., 34 , 415 (1900). 



no THE THEORY CF ELECTROLYTIC DISSOCIATION . 


of I'. This latter change, however, will have to take 
place to only a very slight extent in order to raise the I' 
concentration to the value of the equilibrium, since i A gi 
is very much smaller than L^c \ ; according to Goodwin, 1 
the latter is 1.56 Xio“ 10 and the former 0.94 Xio~ 16 , so 
that the concentration ratio of the Equilibrium (Cl') : (I') 
must be equal to 1600000. From this it follows that 
even with the greatest concentrations of potassium 
chloride, practically speaking, no appreciable quantities 
of potassium iodide are left, but are as good as entirely 
consumed in the conversion of AgCl into Agl. 2 

An exactly analogous case occurring in analytical 
chemistry is Mohr’s method for chlorine titration by 
means of silver solution with chromate addition. Silver 
chromate as well as chloride is difficultly soluble; how- 
ever, the latter much more so than the former, so that the 
equilibrium ratio of the concentrations (Cr 0 4 ") : (Cl') 
is very large. As long, then, as we have present in the 
solution to be titrated much Cl' as compared with the 
small amount of Cr 0 4 " (serving as indicator), essentially 
only AgCl can precipitate upon the addition of Ag* ions. 
Finally, when so much Cl' has thereby been removed 
from solution (practically speaking, all) that -the chromate 
ions can take part in the precipitation, the deep-brown 
silver chromate is formed alongside the white chloride. 
Conversely, in any solution containing an appreciable 
concentration of chloride, the brown precipitate of 


1 1. c. 

2 In case the precipitates form solid solutions with one another, the 
relations .are -changed; -compare F. W. Kuster and Thiel, ^eitgchr, 
anorg. Ch^m-i. 19 , .81;. 23 , 25,. 24 , . 33 v 129 (1.899-1903).. 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. :n 


Ag2Cr0 4 is completely converted into AgCI and soluble 
chromate, a fact employed in sounding ocean depths by 
means of a patent sounding instrument. In this, glass 
tubes are used closed at one end and lined on the inside 
with silver chromate; these, are lowered into the ocean. 
As high as the sea-water enters the tube, corresponding 
to the pressure of the depth, the silver chromate is con- 
verted into white chloride, and the brown-white line of 
demarcation between the two permits the calculation of 
the compression to which the air within the tube was 
subjected at the greatest depth. 

If we have two salts of the same metal, approximately 
equal as to insolubility, in equilibrium with a solution, then 
under all conditions we must also have in the solution an 
approximately equal concentration of the precipitating 
anions, or if they are not equal such equality must be 
brought about by one precipitate being changed into the 
other. 

For example, if we have silver chloride present in a 
solution of silver nitrate, and precipitate the silver ions of 
the nitrate with the aid of KCNS, we obtain AgCNS as a 
second' precipitate. So long as the precipitation of the 
silver is not complete, the solution will not be capable of 
containing sulphocyanate anions; as soon, however, as the 
first excess of sulphocyanate ions is added the equilibrium 
with the AgCI is disturbed, because suddenly the ratio 
(CNS') : (Cl') in the solution is shifted very much in 
favor of CNS'. The equilibrium then adjusts itself in 
such a way that the excess of CNS' ions reacts with AgCI 
to form AgCNS and ’ Cl' ions- until the necessary 
(Cl') : (CNS') concentration ratio is re-established.. The 



H2 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

result Is that a sulphocyanate excess is not shown at once 
by the ferric indicator reaction, and in order to attain 
this, that is, prevent the reaction of the sulphocyanate 
with the AgCl, it is necessary in the Volhard chlorine 
titration method to remove the AgCl by filtration before 
the addition of sulphocyanate. 

The greater the solubility product or the solubility of 
a precipitate is, the greater — for a given concentration of 
one of its ions — the other ion, contained in the u pre- 
cipitating agent,” must become before separation of the 
precipitate sets in. 

The extreme of insolubility is probably that of the 
sulphides, which require as precipitating ions the sulphur 
ions S". These S" ions are contained in greatest con- 
centration in the alkali sulphides, somewhat less in 
ammonium sulphide on account of hydrolysis, very much 
less in hydrogen sulphide, which, according to Walker, 1 
splits up to a just measurable extent into the ions H' and 
HS'. A o. i - NH 2 S solution, which is one almost saturated 
with H 2 S at a pressure of i atm., contains only 0.000075 2 
mole HS' ions per liter; these in turn are further dis- 
sociated to an extremely limited extent according to the 
equation HS'=H‘-fS". Hence the concentration of the 
S" ions in an H 2 S solution is exceedingly small and is 
made still smaller by the addition of acid, whose H* ions 
force down the HS' ions to concentration magnitudes of 
about io~~ 9 . In spite of this, the inconceivably small 
S" concentrations resulting are sufficient to enable a 


1 Zeitschr. physik. Chem., 32 , 137 (1900); Joum. Chem. Soc., 77 , 5 

2 -V^TyXicr 4 *:©.! (see p. 55). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES 113 


whole series of metals to so easily reach their solubility 
products that they are precipitated even from their most 
dilute solutions. 

It is plain to see now why we group qualitative 
analysis according to these sulphide solubilities, for: 

i- The most insoluble sulphides are formed even with 
the few S" ions of a strongly acid hydrogen sulphide 
solution, that is, they do not dissolve in acids (Pb, Ag, Hg, 
Cu, Bi, As, Pt, Au). In other words, they send so few 
S" ions into solution that even with the high H' concen- 
tration of strong acids no H 2 S is produced. 

2. The very insoluble sulphides (Cd, Sn, Sb) are partly 
but not completely precipitated from a strongly acid 
solution, i.e., very small metal ion concentrations no 
longer give a precipitate with the extremely small S" 
concentration, or the sulphides are dissolved (form 
H 2 S) by concentrated acids. 

3. The appreciably soluble sulphides (Zn, Co, Ni) 
precipitate only from neutral or H 2 S solutions acidified 
with a weak acid, but usually not completely until the 
H* concentration of the liberated acid is reduced by 
means of sodium acetate, for example (see p. 71), thereby 
increasing the HS' and S" concentrations. 

4. The markedly soluble sulphides (Mn, Fe) require for 
their precipitation high S" concentration, which is only 
attainable in alkaline solution, i.e., a solution poor in H* 
(ammonium sulphide, sodium sulphide). Even weak 
acids such as acetic acid possess sufficient H' ions to form 
H 2 S with the S" ion of the aqueous solution of Mn or Fe 
sulphide — in other words, to dissolve the sulphide. 

Of interest is also the behaviof of difficultly soluble 



1 14 THE THEORY OF ELECTROLYTIC DISSOCIATION 


oxides and hydroxides, which, in so far as they are soluble 
in water, produce in addition to the particular cathion the 
anions O" and OH' respectively. On account of the 
presence of water the equilibrium condition, (H*) 2 * (O") = 
k and (H*j* (OH'j = £- a , respectively (see p. 55), must 
always be fulfilled. This equilibrium constant is ex- 
tremely small, so the H‘ concentrations of the weakest 
acids are in most cases sufficient to dissolve, with the 
formation of H 2 0 , these difficultly soluble substances. 
Therefore only the most insoluble oxides (and hydroxides) 
are not dissolved in acids to an approximately quantitative 
extent, especially when the acid is at the same time a 
weak one. 

A case of this was found .by Jaeger 1 in dissolving HgO 
in H 2 F 2 - Here he found the ratio of the free acid remain- 
ing in the solution equilibrium to be (H 2 F 2 ) :(HgF 2 ) = 3.6 
as a mean. It is easy to see that this equilibrium is 
nothing more than a hydrolytic one, which is only dis- 
tinguished from the former instances in that the base 
HgO liberated by the action of H 2 0 on the salt HgF 2 
attains its saturation concentration and therefore enters 
equation (15) (p. 90) with constant active mass 


= Hydrolytic Constant 

(bait) (HgF 2 ) 


so that, as found, (H 2 F 2 ) : (HgF 2 ) has a constant value 

( 3 - 6 ) 2 


1 Zeitschr. anorg. Chem., 27 , 26 (1901). 

3 From which by (15), knowing the solubiiiiy 01 Jtigu, tne dissociation 
constant of HgF^ould be calculated. 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . Ii5 

The solution of a precipitate is usually based upon the 
fact that one ion of the substance added unites with one 
ion of the difficultly soluble substance to form an undis- 
sociated body, and thereby disturbs the solubility product 
of the original precipitate. For instance, if we have a 
suspension of BaC0 3 in water, there are present dissolved 
in the water sufficient Ba* * and C0 3 " ions so that we have 

(Ba-):(C 0 3 ")=iBa C03 - 

An addition of acid, H* ions, forms undissociated car- 
bonic acid with the weak CO 3" ions, whereby the CO3" 
concentration is reduced and a corresponding amount 
of the dissolved undissociated BaC0 3 is ionized; then, 
however, the undissociated part ceases to be saturated, 
and consequently more solid BaC0 3 passes into solution. 

In general in the same way every precipitate not too 
insoluble and containing the anion of a very weak acid 
(carbonates, sulphides, cyanides, phosphates, oxalates, etc., 
and especially hydroxides) must be dissolved by H' ions 
(acids), in that the anions are consumed in the production 
of the undissociated weak acids, among which water is to 
be counted. 

Analogously the solvent action of ammonia and its 
derivatives upon the precipitates of many heavy metals is 
explained. The metal ions in these cases are to a very 
great extent taken up by the amines to form complex 
“ amine ” ions, and are thereby removed from participa- 
tion in the solubility product, which in turn strives to 
re-establish itself at the expense of the undissociated 
portion of the dissolved solid, resulting in a solution of 
the precipitate. 



tl6 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

Complex-forming ions, such as cyanogen and iodine 
ions, have a somewhat different kind of solvent action on 
cyanides (Cu, Ag, Cd, Ni, etc.) and iodides (Pb, Hg n , Ag), 
because they take up the undissociated portions of these 
as a “ neutral part ,” 1 * so that the solid must pass into 
solution in order to keep up the concentration of satura- 
tion. 

Special attention has been attracted to several cases, 
more in the nature of curiosities, in which, by the inter- 
action of a difficultly soluble (heavy metal) oxide and 
a neutral salt, alkaline reaction appears, i.e., OH' ions 
are formed. This, for example, takes place between 
HgO and Ell. The dissociation theory also explains this 
phenomenon very simply. Let us consider the oxide 
ion equilibrium, which, since oxides are formed from 
metal M* and OH' ions, may be formulated thus: 

(M-)(OH')=L 0 > 

in which Lq represents the solubility product of the oxide. 
Accordingly, every electrolyte, which consumes the 
cathions M* either for the formation of a difficultly 
soluble compound (as Agl) or for a complex formation 
(as Hgl 4 ", Ag(CN) 2 ', Pbl 3 ', BF 4 '), must bring about an 
increase of the OH' concentration, that is, an alkaline 
reaction. 

So among others the following cases may be predicted, 
and are confirmed experimentally . 3 


1 See Abegg and Bodlander, Zeitschr. anorg. Chem., 20 , 471 (1899). 

3 Heinrich Biltz, in his “ Experimentelle Einfuhrung in die anorgan- 

ische Chemie” (Kiel, 1898), p. 86, maintains that HgO + 2KCN is 




EQUILIBRIA AMONG SEVERAL ELECTROLYTES, n? 


Appearance of Basic Reaction by Interaction of Neutral 
Compounds. 

(Bersch, 1S91 [Ostwald’s Zeitschr., S, 383]; Abegg, 1903.) 


Oxide. 

Neutral Salt. 

Reaction Product. 

PbO 

Potassium 

iodide 

Pbl„ 


C ( 

bromide 

PbBr 0 

Fe(OH), 

1 c 

i C 

fluoride 

FeF 3 (undissoc.) (L) 

1 1 

oxalate 

Fe(C 2 0 4 ) 3 "' (complex) (L) 

Cu(OH) : 

t < 

1 1 

tartrate 

Fehling ion (complex) (L) ‘ 

tt 

sulphocyanate 

Cu-sulphocyanate complex (L) 

< < 

1 1 

thiosulphate 

Cu -thiosulphate complex (ZA 

Ag 2 0 

t i 

i t 

iodide 

Agl 

i £ 

bromide 

AgBr 

1 1 

£ £ 

chloride 

AgCl 

1 1 

£ £ 

thiosulphate 

Ag-thiosulphate complex (L) 

HgO 

£ £ 

iodide 

Hgl 4 " complex ( L ) 

£ £ 

£ C 

bromide 

HgBr/' complex (L) 

it 

£ £ 

chloride 

HgCl 4 " complex (L) 

£ £ 

£ £ 

oxalate 

Hg-oxalate complex (L) 

£ ( 

£ £ 

thiosulphate 

Hg-thiosulphate complex 

Cd(OH). 

« t 

£ c 

£ £ 

£ £ 

£ £ 

iodide 

bromide 

chloride 

1 inner complex, very 
fSSti little dissoc. (L) 

BF 4 ' complex (L) 

B(OH) 3 

£ £ 

fluoride 


All these reaction products, of course, do not react with 
alkalis to separate metal hydroxides. 

A reverse curiosity, the appearance of an acid reaction 
upon the mixing of neutral AgNOg with alkaline 
Na 2 HP0 4 , is explained in an exactly analogous way by 


probably the only reaction of the kind. However, the above con- 
federations of the case, based on the ionic theory, give, as we see, numerous 
reactions. The still more numerous cases in which KCN as a result of 
complex formation brings about alkaline reaction are purposely omitted, 
since, on account of the alkaline reaction of the KCN to begin with, they 
are not as striking as the above produced with absolutely neutral salts. 
In the cases marked (L) in the table, solution of the oxide occurs; in 
the others, conversion into a more difficultly soluble salt- 



US THE THEORY OF ELECTROLYTIC DISSOCIATION. 


the ionic theory. The latter salt contains the ions Na' 
and HP0 4 ", and this anion dissociates with the separation 
of H" ions, according to the equation 

HPO 4 " «=> H' -f P0 4 '". 


This dissociation is, however, exceedingly slight as long 
as appreciable amounts of OH' ions are present, owing 
to the hydrolysis of the Na 2 HP0 4 , which is due to the 
weakness of the anion. Since by the previous equation 
and the law of mass action 

, (HPO/Q 
(PO 4 "0 ’ 



the concentration of the H' ions grows with diminishing 
P0 4 '" ions, and the Ag* ions remove these PO/" ions by 
precipitating Ag 3 P0 4 , so accordingly the addition of 
AgN0 3 must produce acid reaction. 

As for the rest, the appearance of alkaline reaction is 
absolutely analogous to the appearance of acid reaction 
when we use a neutral salt such as CuS0 4 , HgCl 2 , or 
AgN0 3 to remove by means of its cathion (through 
producing insoluble sulphide) the S" ions from H 2 S and 
so force the hydrogen of the same to become ionic. This 
is usually described as decomposition of the salt and 
liberation of its acid by H 2 S. In place of H 2 S, of course 
any weak acid will serve which gives difficultly soluble 
metal compounds, i.e., such whose saturated solutions 
contain very few metal ions. The previously discussed 
case of Na 2 HP0 4 is therefore only a special case of this 
general manifestation. 

The salts of very insoluble hydroxides, such as those 



EQUILIBRIA . AMONG SEVERAL ELECTROLYTES. 1*9 


of Al"’, Cr”', and Fe*", also give an acid reaction with 
the neutral water, because their cathions consume the OFF 
of the latter — a phenomenon that we have already learned 
to recognize as hydrolysis. 

The common feature of all these reactions is that the 
metals and the acid anions combined with OH' and H* 
respectively are contained in considerably smaller 
concentration in the compounds produced (precipitate, 
complex, or undissociated substance). 

Heterogeneous equilibria may also appear in conjunction 
with hydrolysis, as shown by the action of water on tin, 
plumbic, bismuth, antimony, and the strongly dissociated 
mercurous and mercuric salts (nitrate and chlorate). 
Here the OH' ions of the water combine with the very 
weak cathions to form difficultly soluble basic salts, which 
precipitate as soon as hydrolysis has produced such 
quantities of them that they exceed the concentration of 
saturation. As a matter of course, acid reaction (H* ions) 
appears here also and places a limit upon the hydrolysis. 

The hydrolyzing OH' ions may by the addition of acid 
be so reduced at the very outset that no appreciable hydrol- 
ysis and hence no precipitation of basic salt takes place. 
This fact is likewise made use of in analytical chemistry. 

Just as precipitating reactions, and through them the 
heterogeneous equilibria, play a leading part in analytical 
chemistry, so the knowledge of the numerical values 
of the solubility products L 8 for the various precipitates 
is of primary importance. A series of determinations 
of this kind have been made by Goodwin , 1 Immerwahr , 2 

1 Zeitschr. physik. Chem., 13 , 641 (1S94). 

2 Zeitschr. f. Elektrochem., 7 , 477 (1901). 




120 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


Noyes , 1 Kohlrausch and Rose , 2 Bodlander , 3 Sherrill , 4 
v. Ende , 5 Kiister and Thiel, and others , 6 and the following 
figures are taken from these : 


Saturation Concentrations and Solubility Products of 
Difficultly Soluble Salts. 


Ag.O 

Ag- -1.5 Xro- 4 

l 8 = - — 

AglO, 

“ =x-9 Xio~* 

44 =3.6 Xio' 8 

AgCl 

“ =1 .25X io~ 5 

“ = 1.56X10-° (25°) 

AgSCN* 

44 =1.1 Xio~ 8 

“ =1.2 Xio- 13 

AgBr 

44 =6.6 Xio -7 

“ =4*35X IO “ t3 (25°) 

Agl 


44 =1.0 Xio- 16 * 4 

Ag 2 Cr 0 4 

14 = 1.7 Xxo- 4 

“ =1.0 Xio- 11 (iS c ) 

TICi 


44 = 2 . 65 X IO— 4 (25°) 

TIBr 

44 =2.0 XlO“ 3 

44 =4.0 Xio* 8 

TIBr 

“ =8.7 Xio- 3 

44 =7.6 X io~ 5 (68-5° j 

TISCN ....... 

44 =1.5 Xro~ 2 

44 = 2 . 25 X 10 4 (25°) 

tlso 4 ....... 

44 =9.0 Xic~ 2 

44 =3-6 Xio~ 4 

CuCi 

.. .. Cu* =1.1 Xio -3 

“ =1.2 xio-° 

CuBr 

“ =2.0 Xio- 4 

“ =415x10-* 

Cul 

“ =2. 25X10-' 

“ =5.1 Xio- 13 

HgjClj 


“ = 3-5 Xio- ,s (25 c ) 

HgJJi-j 

' 44 =7-0 Xic -8 

“ =1-3 Xio -21 

Hg,J„ 

“ =3.0 Xio- 10 

“ =1.2 Xio -3 ’ 

Hg^o 4 

44 =8.^ Xio- 4 

“ =3.0 Xio— 0 

HgCI, 


“ =2.6 Xio- 15 

HgBr, 

■ 44 =2.7 Xio~ 7 

“ =8.0 Xio- 30 

Hgl 2 

...... “ =2.0 Xio- 10 

“ =3-2 Xio- 3 * 

PbCIj 

Pb” =2.0 X io~ 2 

44 =1.0 Xio -4 

PbBr 2 

44 =2.0 Xio -2 

44 =6.0 Xio- 8 

Pbl 2 

“ =i.S Xio- 3 

44 =1.0 Xio -7 

PbSO. t 

..... 44 =1.5 Xio- 4 

44 =2.2 Xio -8 


1 Zeitschr. physik. Chem., 6, 241 (1890); 42 , 336 (1903). 

2 Ibid., 12 , 241 (1893^ 

3 Zeitschr. anorg. Chem., 31 , 474 (1902). 

4 Zeitschr. physik. Chem., 43 , 705 (1903). 

5 Zeitschr. anorg. Chem., 26 , 129 (1901). 

8 Ibid., 24 , 57 (1900), and 33 , 129 (1903); see also Wilsmore, Zeitschr. 

physik. Chem., 35 , 305 (1900). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 121 


ANOMALY OF STRONG ELECTROLYTES. 


While in the preceding discussions it is safe to say the 
law of mass action, as applied to an extended series of 
reactions between ions and undissociated substances, 
found its excellent verification, yet it seems in one es- 
pecially simple and important case to utterly fail, namely, 
in the dissociation of the so-called strong electrolytes — 
the salts — and the strong acids and bases. Whether we 
derive the degree of dissociation of these electrolytes from 


the conductivity 



or from the freezing-point 



, we arrive at values for the expression 


C£ 2 ‘C 

of 

i —a 


the law of mass action which for the different concentra- 
tions c decidedly deviate from the demanded constant, 
and therewith prove that, for some reason, either the law 
of mass action is to be modified for these electrolytes, or 
the methods for the determination of the degree of dis- 
sociation a in these cases give incorrect values. 

Recently this vulnerable spot of the dissociation theory 
has been very assiduously investigated and discussed, and 
we are indebted to Jahn in particular for a series of 
brilliant measurements of precision. He attempted to 
get at the degree of dissociation by a third method, namely, 
that of the measurement of concentration chains. Up 
to this time this method had hardly been used, because 
of its lack of sensitiveness for purposes of ordinary 
accuracy. 

It would take us too far to give here the theoretical 
considerations which rest on the exact application of 



122 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


thermodynamics, especially as up to the present we can 
by no means look upon the problem as solved. There 
is a tendency on the one hand to assume that in the 
solutions of strong electrolytes the ratio of the equivalent 
conductivities for different concentrations cannot give the 
degree of dissociation accurately, because, in consequence 
of the variable friction, the mobility of the ions varies in 
the solutions according to the amount of salt contained 
and cannot be assumed as equal. On the other hand, 
the osmotic methods (depression of the freezing-point, 
etc.) might fail, for reasons which may be of a physical 
as well as of a chemical nature. 

Nemst and Jahn 1 find the physical reasons in that 
there exists in the solution an interaction between the ions 
and the undissociated molecules, which counteracts their 
mutual independence and so causes the osmotic pressure 
to be different from that which, according to van’t Hoff’s 
law r , corresponds to the concentration. 

From the fact that the osmotic laws hold for non- 
electrolytes and weak electrolytes, Jahn draws the con- 
clusion that such an interaction between the undissociated 
molecules may be neglected. He further makes it 
plausible that the ions have no marked influence on each 
other, since the electrostatic attractions of the unlike- 
charged are just counteracted by the repulsions of the 
like-charged; that is, there 'would be left only the inter- 
action of the undissociated molecules with the ions, which 
would have to be considered accountable for the devia- 


1 Zeitschr. physik. Chem., 33 , 545, 35 , 1; 37 , 490; 38 , 487; 41 , 257 
Q1900-1902). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 123 


tion of the osmotic pressure from vairt Hoff’s law. The 
mathematical form in which this is to take place has been 
shown in the latest investigation of Jahn, just referred to. 
The result is that the dissociation constant is not given 
by the formula 

9 9 0(211— N) 

n z n- — 

= £, but should be expressed by e K ° =k, 

N—n r J N —n ’ 

in which N is the total molecular concentration, n the 
ionic concentration, no the molecular concentration of the 
solvent, e the base of natural logarithms, and a the 
characteristic constant of the interaction named. 

That this formula agrees well with the freezing-point 
determinations is shown by the following calculation of 
Jahn, using Abegg’s freezing-point determinations on 
KC1 with the selection of a suitable a value: 


N . . 0.0237 0.0354 0.0469 0.0583 0.0697 

n c.0208 0.0302 0.0384 0.0463 0.0525 

2 a( an—N) 

— e n ° ..0.132 0.147 0-141 0.141 0.125 Mean: 0.137 


The applicability of Jahn’s equation only shown, how- 
ever, as Jahn himself states, that the physical explanation 
is correct as to formulation. Indeed, it seems to us that 
a number of considerations demand another interpreta- 
tion of these complicated relations. The physical point 
of view should lead to the conclusion that the interaction* 
between ions and molecules should manifest itself with all 
strongly dissociated electrolytes; at least, it would not be 
dear why this influence should assume markedly different 
values for substances of a similar degree of dissociation. 



124 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


But after all among Ostwald s 1 extended material there 
are to be found a number of strongly dissociated acids, 
such as dichloracctic acid, maleic acid, cyanacetic acid, 
and various bromine-substituted amidobenzene-sulphonic 
acids, for which Ostwald' s simple dilution law, even up to 
degrees of dissociation as high as 98%, gives good con- 
stants. The following summarization contains some 
figures pertaining thereto; under 100a we have the 
degrees of dissociation in percentages up to those for 
which the dissociation constant k holds, and under v the 
dilutions for which the value 100a: holds: 


Strong Electrolytes which Obey the Dilution Law. 



TO ? k 

JO 0 a 

V 

CL»-acetic acid 

5-14 

93-4 

256 

CN-acetic acid 

0.37 

>82.1 

1024 

Maleic acid 

1. 17 

92.8 

1024 

<?-NHo-benzene-sulphonic acid 

<=>■33 

>80 

1024 

(1:2:5) Br-NH 2 -benzene-sulphonic acid. . 

1.67 

97 

1652 

(1:2:415) Br 2 -NH 2 -benzene-sulphonic acic 

7-9 

97.8 

556 

(1:3:4: 5) Br 2 -NH 2 -benzene-sulphonic acid 

2 -5 

96 

115° 


Again, other acids show the behavior characteristic of 
strong electrolytes in spite of great analogy in composition. 
In any case, according to this, it is not very likely that the 
high degree of dissociation essentially determines the 
deviation from the law of mass action, and certainly in 
the cases of the above-tabulated acids, at least, the con- 
ductivity is to be looked upon as a correct measure of 
the degree of dissociation. 


1 1. c.; see p. 29. 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 125 

Further, the physical assumption of Jahn still owes us 
an explanation as to whether or not the interaction which 
exists between many ions and few molecules (strongly 
dissociated substances) is not likewise observable between 
many molecules and few ions (weak electrolytes), as 
every analogy would lead us to expect. 

The chemical facts which come into consideration for 
an explanation of the anomaly are on the one hand the 
formation of inner complexes, and on the other the 
hydration of the ions. 

The formation of inner complexes was discovered by 
Hittorf 1 on cadmium salts in their transference behavior 
during electrolysis, and since then has been accepted 
as the explanation for the variability of the transference 
number with the dilution. It consists in an addition of 
the undissociated molecules to one of the ions of the 
electrolyte, and the extent to which it takes place depends 
on the concentration of the two components of the complex 
(ion and undissociated part). A quantitative investiga- 
tion of inner-complex formation has not been possible 
thus far, but in addition to the investigations of Hittorf, 
those of Bredig , 2 Noyes , 3 and especially those of Steele 4 
should be mentioned, which have experimentally placed 
the fact beyond doubt. Of these, Steele especially 
discusses in detail the necessity of this assumption of 
Hittorf. 


1 Pogg. Ann., 106 , 385 and 546 (1859). 

2 Zeitschr. physik. Chem., 13 , 262 (1894). 

*Ibid., 36 , 63 (1901). 

4 Ibid., 40 , 722 (1902). 


226 the theory of electrolytic dissociation. 

That inner-complex formation is present even in the 
salts of very positive metals is made highly probable by the 
existence of such double salts as K 3 Na(S04)2 (glaserite), 
KMgCl 3 (camallite), etc. We shall therefore have to 
take into consideration, even in the case of salts such as 
NaCl, etc., the possibility of an ionic formation such as, 
say, Na* and NaC^'- With salts of metals of less electro- 
affinity, ionic formations of that kind are beyond doubt, 
as was experimentally demonstrated in every direction 
for C0CI2, CuCl 2 , ZnCl2, etc., in the nice research of 
Dorman, Bassett, and Fox . 1 

It is well also to call attention to the fact that in the 
exact equation as developed by Jahn 2 for the electromotive 
force of concentration chains there is, in addition to the 
logarithmic factor with the ratio of the ionic concentra- 
tions, another factor proportional to the difference of the 
ionic concentrations. Such a factor would have to be 
present in case of the existence of ionic hydrates in order 
to give due consideration to the water combined with the 
ions 3 in the work of their transport from one concen- 
tration to the other. 

The formation of inner complexes would reduce the 
concentration of the independent molecules so that the 
osmotic methods (freezing-point, etc.) would give smaller 
i values, as well as smaller a values, than demanded by 


1 Trans. Chem. Soc., 81 , 944 (1902). 

2 Zeitschr. physik- Chem., 41 , 276 (1902). 

3 Compare Dolezalek, Theorie des Bleiakkumulators, p. 35 (Halle. 
igoi); also translation. Theory of the Lead Accumulator, p. 65 (Wiley 
& Sons, 1904); and F. Haber, Zeitschr. physik. Chem., 41 399 (1902). 




EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 127 

the law of mass action; accordingly, the dissociation 
constants, as calculated by means of the freezing-points, 
should diminish with increasing concentration. This 
agrees, for example, with the experience 1 on RbN 0 3 , 
though usually one observes in the constant thus calcu- 
lated decidedly the reverse course. 

Just as with their own undissociated molecules, the 
ions also form complexes with the solvent (hydrates), 
whose existence is likewise supported by extended ex- 
perimental evidence. 2 Such a formation of hydrates, 
in distinction from the formation of inner complexes, 
would leave the number of moles of the dissolved electrolyte 
unchanged, while that of the solvent would be diminished, 
and so with increasing concentration one w r ould reach 
an accelerated increase of the molecular concentration 
ratio, electrolyte : water, which agrees qualitatively w ith 
the course of the dissociation constants calculated from 
the freezing-points. 3 Even if the quantitative foundation 
for this explanation is wanting, it is still noteworthy and 
seems to speak in favor of the chemical explanation that 
csesium nitrate, a salt of whose ions, according to a theory 
of Abegg and Bodlander (see later on p. 161), one is led 
to expect a minimum tendency to form complexes and 
hydrates, gives degrees of dissociation according to its 
freezing-point depressions,, as determined by W. Biltz, 4 
which are entirely in accord with the law of mass action, 
as the following table shows : 


1 W. Biltz, Zeitschr. physik. Chem., 40 , 217 (1902). 

2 For literature, see Biltz, 1 . c., p. 214. 

3 For examples, see Jahn, 1 . c. 

4 1, c.j p. 218, 



128 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


Ci£3lU:.I XlTSATE. 



J 

j 




c 

1 X . 85c 


( 

i-«) 

0.00766 

0.0194 

0.0465 

0.098S 

0.142 

0.210 

O.299 

O.386 

0.434 

3.66 

3.6 1 

3-53 

3-35 

3 - 2 4 

3 - r 5 

3-°37 

2.914 

2.92 

1.98 

i -95 

1. 91 

1. 81 

i -75 ! 

1 . 704 
1.64 
i -575 
1-578 

0.9S 

°-95 

0.91 

0.81 

°- 75 

0.704 

0.64 

0 - 5 75 
0.57S 

c- 33 " 

o -35 

0.41 

0-34 

0.32 

o -35 

c -34 

c - 3 ° 

0.34. 

- Mean: 0.34 


Since the degrees of dissociation taken from the con- 
ductivity (^ a== 2 ~J °f caesium nitrate by W. Biltz and Jul. 

Meyer gave markedly greater values than the above and 

a?c 

also led to no constant for ;__ a y ^ P^i 11 to see that 

the conductivity in the case of this and no doubt many 
other salts is not a correct measure of the dissociation, 
while before (see p. 124), in regard to a number of acids 
of moderate strength, we were forced to the opposite con- 
clusion. 

With the anomalous strong electrolytes it seems that the 
conductivity in almost every instance gives the semblance 
of too high degrees of dissociation. Thus, for example, 
according to Biltz and Meyer, CsN 0 3 gives 

for c =0-25 0.125^ 

Freezing-point a = 0.67 0.78 

Conductivity a = 0.76 _ 0.82 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES. 129 


that is, with growing concentration we have a very marked 
increase in the discrepancy. This may, as Jahn assumes, 
be due to a decreasing ionic friction in the more con- 
centrated solutions; however, several conclusions follow- 
ing from this are not, as tested by Sackur , 1 borne out 
by experiment. 

We see, then, that this material still demands extended 
and searching investigation, but so much at least it seems 
we may say at present with reasonable certainty, that the 
law of mass action will prove itself, as usual, to hold 
absolutely also for strong electrolytes, as soon as we come 
into possession of perfect methods for obtaining the real 
ionic concentrations or degrees of dissociation. 

As to strong electrolytes we must console ourselves, 
in so far as a complete molecular theoretical explanation 
is concerned, with the hope of a possibly near future; 
however, let us in addition mention several attempts to 
express the course of their conductivities in a mathemat- 
ical formula. Rudolphi 2 was the first to give an equation 
for this, which was later transformed by van’t Hoff and 
Kohlrausch 3 into 

(1 — a ) 2 

in which, as hitherto, a = -j-. The physical significance 
of this formula is: 

(Ionic concentration) 3 ^* (Undissoc.) 2 . 

1 Zeitschr. f. Elektrochem., 7 , 475 (1901). 

3 Zdtschr. physik. Chem., 17 , 385 (1895). 

•Ibid., 18 , 301, 662 (1895). 


130 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

In order to give an idea of the extent to which this equation 
conforms to the observations, the following small table 
may be offered: 


Constants of van’t Hoff’s Dilution Law for Strong Binary 
Electrolytes. 


(van’t Hoff, 1S95.) 


V 

! kno 3 ! 

| (is°) : 

MgS 0 4 

( 18 0 ) 

HC 1 

(18°) 

KC 1 
(99- 4 °) 

KCl 

(18°) 

NaCl 

(18°) 

KBr 

(18 0 ) 

V 

LiCl 

(18 0 ) 

2 i 

1.63 

— 

4.41 

1.83 

2.49 

1.87 

2.44 

2 

1.27 

4 

1.67 

0. 162 

4.87 

1 -79 

2.2 3 

I. 71 

2-55 

IO 

1. 16 

8 

1.68 

0.156 

4-43 

1.76 

2.1 

1.6 

2.28 

. 20 

1.07 

16 

1 . 72 

0.151 

4.72 

1.92 

I -94 

i -4 

2. 3 S 

33-3 

1.02 

3 2 

1.82 

0. 151 

5- 2 9 

i -9 

1.87 

i -43 

2.4I 

100 

0.92 

64 

i 

i.SS 

1 0.158 


1.78 

1.72 

1.38 

2 . 72 




For the sake of completeness we must mention an in- 
vestigation by Storch, 1 who assumed that the dilution law 
is represented by a formula in which any power of the 
ionic concentration is written proportional to any other 
power of the concentration of the undissociated portion. 
In the formula of van’t Hoff the ratio of these powers is 
3:2, that is, equal to 1.5. Storch found by extended 
calculations that the power ratios, which are reproduced 
as nearly as possible in the conductivity observations, are 
somewhat different for the different electrolytes, KC 1 , KI , 
KOH, KNO s , HC 1 , HN0 3 , MgS0 4 , CuS0 4 , H 2 S0 4 , : 
K 2 S0 4 , BaCl 2 , ZnCl 2 , varying, however, only between 
1.40 and 1.577, so that in any case they approach very 
closely the value 1.5 'of* the van’t Hoff formula. The 
close agreement with observed values makes it possible 


1 Zeitschr. physik. Chem., 19 , 13 (1896), 


EQUILIBRIA AMONG SEVERAL ELECTROLYTES . 131 

to foresee that the formula, 1 for the present purely em- 
pirical, will some day give an interesting relationship 

A 

between the real degree of dissociation and the -7- values. 

A 0 

Before dismissing the question of the variability of the 
degree of dissociation with the concentration, it is neces- 
sary to inform ourselves as to the concentrations up to 
which the equation of the law of mass action or the so- 
called dissociation isotherm does hold. We may say, a 
priori, that it cannot hold without limits, for if we 
Continuously increase the concentration of the electrolyte, 
in the first place we depart from the field of Cl ideal 
dilute ” solutions, in which alone the van’t Hoff laws 
for the osmotic pressure hold strictly, and secondly, the 
replacing of the solvent by dissolved electrolyte alters 
more and more the medium in which the equilibrium has 
to establish itself. We can no more expect of this reaction 
than of any other that it be independent of the nature of 
the solvent.. The same conclusion follows from the 
consideration that in so far as we are dealing with dis- 
solved electrolytic liquids for the limiting case of highest 
concentration, i.e., water- free electrolytes, the concentra- 
tion of the ions must be extremely small, since even pure 
liquids, according to a rule to be mentioned later (see p. 
155), possess only a very small conductivity of their own. 


1 The recent attempt of Roloff (Zeitschr. angew. Chem., 1902, Heft 
22-24, also separately published by Springer, Berlin, 1902) to establish 
a theoretical basis on the ground of the assumption of a variable disso- 
ciation of the water itself stands in direct contradiction to the facts, 
for the ionic product of water, even in concentrated solutions of electro- 
lytes, has been found by the most "diverse methods (see p. 56) to be 
entirely constant. 



*3 2 the theory of electrolytic dissociation . 

The held of these more concentrated solutions has 
recently been the subject of interesting researches by 
Wolf 1 and Rudorf, 2 whose results may best be dis- 
cussed with the aid of one of the examples studied, 
namely, that of acetic acid. It w ? as shown that the 
addition of acetic acid to any electrolyte acts in such a 
way as to reduce the mobility of the ions of the latter by 
a definite amount, proportional to the concentration of the 
acetic acid. This reduction of mobility is equal to about 
9-3% P er normal strength of the acid, so that, for instance, 
in a mixture of sodium chloride -hi -normal acetic acid 
the conductivity of the sodium chloride, no matter in 
what concentration it is present, is 9.3% less than in pure 
water. Since this influence upon the mobility of ions has 
shown itself to be entirely independent of their nature, 
nothing is easier than to assume that the acetic acid also 
affects the mobility of its own ions in the same measure, 
so that in a 1 normal acetic acid, for example, the value 
of the equivalent conductivity must again be less by 
9.3% than if the same number of ions moved in pure 
w'ater, as is the case at infinite dilution (A 0 ). In order, 
therefore, to get at the real degree of dissociation, the 
equivalent conductivity A for the high concentration is to 
be increased by the correction due to the retarding in- 
fluence which the presence of the undissociated acid 
brings about. In fact, Abegg found that the constancy 
of the expression 

(Corr. J) 2 -c ^ 

A 0 -(A 0 ~Corr.A) = ConSt - 

1 Zeitschr . physik . Chem ., 40 , 253 ( 1902 ).. 

2 Ibid ., 43 , 257 ( 1903 ). 



EQUILIBRIA AMONG SEVERAL ELECTROLYTES . *33 


is fulfilled up to considerably higher values for c than is 
the case when the correction is not introduced. The 
following values are taken from the investigation of 
Rudorf, who took up the subject more exactly: 


Acetic Acid (25 0 ). ^ 0 = 397 - 


c 

A 

k (uncorr.) 

Corr.-factor. 

k (corr.) 

0.019 

12.3 

1.88 

1 . 002 

i. 88 

0.039 

8-57 

1.85 

1.004 

1.87 

0.078 

6.08 

1 . 84 

1. 007 

1.87 

°* I 57 

4-23 

1. 81 

1.015 

1.86 

°- 3 T 3 

2.97 

1.76 

1.C29 

1.86 

0.627 

2.01 

1 . 62 

1.059 

1.82 

1-254 

I.29 

i -33 ! 

1. 11 8 

r.67 


The range of the reaction constants, or in other words 
the field within which the affinity manifestations of the 
ions possess the same intensity as in pure water or at 
extreme dilution, extends in this case to about o. 6-normal 
concentration. At higher concentrations where the con- 
stant, even by introducing the corrected A value, assumes 
other values than for dilute solutions, this deviation is 
undoubtedly to be attributed to the changing of the 
medium. And indeed I may draw the conclusion from 
Rudorfis investigation just alluded to, that this constant, 
variable with the medium, possesses the chemical signifi- 
cance belonging to it according to its derivation; for the 
same also regulates the ratios of the quantity of acetic 
acid ions and the undissociated acid, when the acetate 
ions are altered by the addition of strongly dissociated 
acetates. 



INFLUENCE OF PRESSURE AND TEMPERATURE 
ON DISSOCIATION. 


Now that we have considered in the preceding pages the 
influence of concentration on the degree of dissociation of 
electrolytes, the question arises, Upon what other addi- 
tional influences is dissociation dependent? A general 
answer is given by the so-called thermodynamic principle 
of Le Chatelier, according to which, action on a system 
in equilibrium by an agent from without brings about 
a reaction that works against this external action. If, 
then, we attempt to increase by compression the pressure 
on such a system, that one of the two reactions in equi- 
librium (decomposition into ions or combination of ions to 
form undissociated molecules) will take place which gives 
a diminution in volume, for thereby the condition of 
stress in the system caused by the pressure is reduced. At 
the suggestion of Arrhenius, a fine piece of research 
pertaining to this influence was carried out by the Russian 
Fanjung, 1 w T hom, we regret to say, death claimed all too 
early. He worked with pressures as high as 260 atmos- 
pheres. The observed increases in conductivity for the 
highest pressures amount at most to about 9%, depending 
upon how great is the volume-difference between the 


1 Zeitschr. physik. Chem., 14 , 673 (1894). 




INFLUENCE OF PRESSURE AND TEMPERATURE . i$5 


undissociated and dissociated acids (mostly organic). 
They are found to be in best agreement with the values 
calculated from this volume-difference. It is plain to 
see, however, that the pressure variations occurring in 
every-day life have no effect in any way noteworthy on 
the dissociation of electrolytes. 

On the other hand, the influence that temperature 
changes may have on dissociation is very marked. If we 
wish to determine this from a study of the conductivities, 
it is necessary, in the first place, to consider the super- 
imposed influence of the changed ionic friction or ionic 
mobility discussed on p. 35 and follow- ing pages. The 
conductivity changes conditioned on mobility evidently 
are not linked writh a temperature influence on the degree 
of dissociation, for the variation in conductivity caused 
by variation in temperature can be attributed, and in fact 
in many cases is to be attributed, essentially to changed 
ionic mobility, without the degree of dissociation of the 
electrolyte having at the same time undergone any altera- 
tion. To put it mathematically, in the expression 

a=~* not only A but Aq as well is changed in the same 

ratio by the temperature. 

Again, the temperature influence on the dissociation is 
given by applying the principle of Le Chatelier. The 
addition of heat will alter the degree of dissociation in the 
direction of that reaction which absorbs heat; that is, in 
case the ionic dissociation is endothermic the dissociation 
will increase upon heating; in case it is exothermic it will 
decrease, A number of methods may be employed to 
determine the heat effect of ionic decomposition or the 



T 36 THE theory of electrolytic dissociation. 


heat of dissociation (ionization), which are based on the 
previously mentioned law of the thermo-neutrality of 
strong electrolytes. According to this law, the mixing of 
two electrolytes produces no heat effect when all of their 
constituents continue in the ionic state after mixing. 
Take, for example, the electrolytes KC2H3O2 and HC 1 , 
and let us assume, contrary to the facts, that upon mixing 
they do not form imdissociated acetic acid, but that the 
acetate' and H* ions continue to exist alongside of each 
other; then in this case the mixing would not involve a 
heat of reaction. However, the heat of reaction occurring 
in reality is to be attributed directly to the circumstance 
that H* and acetate' ions unite to produce undissociated 
acetic acid, and so we have in this heat of reaction be- 
tween potassium acetate and hydrochloric acid the imme- 
diate heat effect of the formation of undissociated acetic 
acid from its ions. The negative value would therefore 
represent the heat of dissociation of acetic acid. This 
method of determining the heat of dissociation may be 
described as the mixing of two strongly dissociated com- 
pounds, of which each contains one of the ionic com- 
ponents of the weakly dissociated substance whose heat 
of dissociation is sought, and which in the process of 
mixing is formed in the undissociated state. 

Another method, very similar in principle to the one 
just mentioned, is based on the fact alluded to above 
(see p. 59), that the neutralization of strong acids and 
bases by one another always produces the same heat effect, 
13700 cal. per gram-equivalent. Deviations from this 
heat of neutralization, -which essentially represents the 
heat effect of the formation of water from H* and OH 7 



INFLUENCE OF PRESSURE AND TEMPERATURE . 137 


ions, are to be found whenever weak acids or bases are 
neutralized by one another, as the table below shoTvs. 

The deviations explain themselves in that, in addition to 
the formation of water, a further reaction takes place. Since 
the neutral salt solution formed is strongly dissociated 
according to the general rule, while before neutralization 
one of its ionic components was appreciably undissociated, 
being combined with one of the ions of water, it follows 
that in such a neutralization a dissociation of the weak 
electrolyte employed must result at the same time. The 
equation of such a reaction will best elucidate the facts. 
For that purpose let us consider, say, the neutralization of 
NaOH by acetic acid, and assume as an approximation 
that NaOH is completely dissociated, while only the 


Heats of Neutralization. 
(Thomsen.) 


By NaOH. 

Weak Acids. 

Metaphosphoric acid . . 14300 cal. 

Hypophosphorous acid. 15 100 “ 

Hydrofluoric acid 16300 tc 

Acetic acid 13400 “ 

Chloracetic acid 1430° “ 

Dichloracetic acid 14800 11 

Strong Acids . 

HC 1 13700 cal. 

HBr 13700 “ 

HCIO3 13800 “ 

HN 0 3 13700 “ 


By HC 1 . 
Weak Bases . 


Ammonia 12200 cal. 

Methylamine 13100 tc 

Dimethylamine 11800 <c 

Trimethylamine. ..... 8700 * 4 


Strong Bases . 

LiOH 13800 cal. 

NaOH 13700 “ 

Ba(OH) 2 13900 “ 

T etram ethy lammonium 

hydroxide 13700 “ 


(small) fraction a per mole of the acetic acid is present 
in the form of ions, and the larger portion (1 —a) is in 
undissociated combination with the H* ions. If now we 



13 s THE THEORY OF ELECTROLYTIC DISSOCIATION . 

mix one mole of each of the two substances and write 
the equation in such a way that we separate the ions and 
undissociated portions from each other, then the neutrali- 
zation equation will read: 

i Na*-f i OH' -Fa H'4- a Acetate' 4 - (i — a)H acetate 

= i Na* 4 -i Acetate' 4 - H 2 0 +a cal. 

If we combine with it the further assumed equation 

i N a* 4 - 1 OH' 4- 1 H‘ 4 -i Acetate' 

= i Na* 4 -i Acetate' 4 - H2O 4 - 1 3 700 cal., 

which would apply if acetic acid were a strong, almost 
completely dissociated acid, we get, by subtracting the 
first equation from the second, the following simple 
expression : 

(1 — a) H ' 4- (1 — a) Acetate' 

= (1 —a) H acetate 4- (13700 — a) cal. 
or 

H acetate =H* 4 - Acetate ' 4 — cal. 

1 —a 

That is, the observed heat of reaction a diminished by 
13700 cal. represents the heat of dissociation for i—a 
moles of the weak acid or base when we neutralize the 
same by a strong base or acid and divide the heat effect, 
diminished by 13700 cal., by i—a . With a weak acid 
1 — a=i , very nearly, a being very small. 

Experimentally this way of obtaining the heat of dis- 
sociation is not very advantageous, because it generally 
gives the sought magnitude as a small difference of two 



INFLUENCE OF PRESSURE AND TEMPERATURE . 139 


large heat effects, so that the experimental errors play 
a very important part in this difference. It is unlike the 
previous method in that we start with the weakly dis- 
sociated compound and end with the strongly dissociated 
salt of the same; that is, to a certain extent it is the 
reverse of the first, in which we prepare the undissociated 
substance from the strongly dissociated salt by means of 
a strongly dissociated acid or base. The heats of disso- 
ciation obtained by either of these two methods may 
serve to calculate the influence of temperature change on 
the degree of dissociation by employing the thermodynamic 
equation derived by Arrhenius, 1 

dink W 
dT ~~ RT 2 9 

in which k signifies the dissociation constant, W the heat 
of dissociation, T the absolute temperature, and R the 
gas constant in calorimetric units (1.99). 

It is true Arrhenius followed the reverse course, in that 
he calculated the heat of dissociation W from the vari- 
ability with the temperature of the conductivity or of the 
dissociation constant k, and compared it with the results 
obtained by the methods indicated above. From these 
conceptions of Arrhenius it was possible to predict an 
interesting case of conductivity: that with electrolytes 
whose heat of dissociation is strongly positive,- rise in 
temperature is followed by such marked reduction in the 
degree of dissociation that the increase in conductivity 
resulting from the enhanced ionic mobility is covered up. 
In fact, in this way it has been possible, in cases like 


1 Zeitschr. phy$ik, Cbem., 9 , 339 (1892). 



140 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

phosphoric and hypophosphorous acids, to establish 
experimentally such reductions in conductivity. This 
was the more striking since it was generally thought that 
the characteristic difference between electrolytic and 
metallic conductivity was the negative temperature 
coefficient of the latter, i.e., reduction in conductivity 
with rise in temperature. 

In the following table a series of heats of dissociation is 
given, from a consideration of which, however, nothing 
in the way of a relationship between these values and the 
chemical nature of the substances has resulted. It is 
worthy of note nevertheless that they are all markedly 
variable with the temperature. 

Heats of Dissociation. 

(Arrhenius, 1SS9, 1S92; Thomsen, 1SS2; Baur, 1897.) 

(Exothermic dissociation taken negative, as is customary in 
thermodynamics.) 

35 ° 21.5 0 


Acetic acid — 386 cal. +28 cal 

Propionic acid — 557 “ — 183 44 

Butyric acid — 935 44 — 427 44 

Succinic acid 4- 445 “ +1115“ 

Dichloracetic acid — 2893 4 ‘ — 2924 4 * 

Phosphoric acid — 2458 44 — 2103 44 

Hypophosphorous acid — 4301 * * — 3745 * 4 

Hydrofluoric acid — 3549 * * . — 

Water (xo. 14 0 ) 4- 14247 44 ■ (24.6°) + 13627 44 


Interpolation-formula: - °° after Kohlrausch and Hevdweiller.* 

(273+/) 


i 

1 5° 

r 5° 

25° 

35° 

Nitrourea 

•••i +5477 

4-3812 

+3640 



Nitrourethane 

+3665 

+ 3724 

+ 2943 

4-2260 

Amidotetrazole .... 

-•* T-J724 

+5 2 5 s 

-1-4593 

+3865 



INFLUENCE OF PRESSURE AND TEMPERATURE . I 4 1 


The heat effects are in part positive, in part negative, and 
in many instances quite small, which is in keeping with 
the fact that frequently the dissociation into ions takes 
place without very great energy changes. The heats of 
dissociation of salts as calculated by Arrhenius are to be 
looked upon as uncertain, for the reason that (see p. 121) 
we are still in doubt on the dissociation of the same as 
calculated from the conductivities. The substances of 
special interest are those with very great heat of dissocia- 
tion, because this corresponds to a great variation of the 
degree of dissociation with the temperature, as we saw 
above from the conductivities of the two acids of phos- 
phorus. The most interesting substance in this regard, 
because the most extreme, is water, whose decomposition 
into ions absorbs the enormous quantity of 13700 cal. of 
heat. From this, according to the equation of Arrhenius, 
we can predict that the usually small temperature in- 
fluence on the dissociation, and therewith on the con- 
ductivity, must be abnormally great in the case of water. 
Kohlrausch and Heydweiller, in their previously men- 
tioned investigation, had occasion to test this conclusion 
of the dissociation theory. They calculated the tem- 
perature coefficient of the conductivity of pure water by 
the equation of Arrhenius and found it to be 5.8% per 
1 0 , 1 while the otherwise largest known temperature co- 
efficients are those of the salts, which at most scarcely 
amount to one hah as much. The investigators named 
found the highest value for the temperature influence on 
the conductivity of water, that is, the influence peculiar to 


1 Wied. Ana., 53 ,. 231 (1894). 



142 THE THEORY OF ELECTROLYTIC DISSOCIATION . 

water least contaminated, to be 5.3%. and they were able 
to calculate, by means of the deviation of this value from 
that theoretically found, that the purest water obtained 
by them still contained impurities to the extent of several 
thousandths of a milligram per liter. For the pure water 
the observations gave conductivities at the different tem- 
peratures which led to the degrees of dissociation con- 
tained in the following table: 


Dissociation of Water at Different Temperatures. 
(Kohlrausch and Heydweiller, 1S94.) 



O) 

0 

I O 0 j lS° ; 20° j 

34 ° 

42 0 

50 ° 

roo° 


Ionic concentra- 
tion/Liter. . . 

k w ! 

I 1 

i c * 35°*39 

O. 120. 15 

i i | 

A 0 i i 

0.56:0.80 1.09: 
O.3TI0.64 1.2 

! .■ j 

1 - 47 

2- 15 

i -93 

3-7 

2.48 

6.15 

8.5 

72.O 

xio-’ 

Xio- u 

i 


We see how rapidly the dissociation of the water rises 
with the temperature, and that, for example, at 50° it is 
already more than three times greater than at the ordinary 
temperature. For the phenomenon of hydrolysis this 
fact is of very great importance, since, as we saw (p. 76), 
the degree of hydrolysis is determined by the ionic con- 
centration raised to the second power, which is the water 
constant k w . Correspondingly, hydrolysis greatly increases 
at higher temperatures. A whole series of chemical 
experiences . may be explained on this basis. If, for 
example, we color a neutral ammonium salt solution with 
litmus and heat it, w r e observe that at higher temperatures 
a distinct red color sets in, indicative of the fact that an 
‘appreciable quantity of free hydrochloric acid has been 



INFLUENCE OF PRESSURE AND TEMPERATURE 143 


separated, or, in the language of the dissociation theory, 
undissociated ammonium hydroxide has been formed 
from the ammonium ions and the hydroxyl ions of the 
water. A further phenomenon belonging here and 
made use of in analysis is the precipitation of difficultly 
soluble hydroxides by inducing the hydrolysis of their 
weak salts. This happens, for instance, in the precipitation 
of basic ferric acetate from a solution of ferric chloride by 
an alkali-acetate, and it may be of interest to consider this 
action somewhat more in detail. These last two salts 
we may look upon as markedly dissociated, notwithstand- 
ing the fact that the ferric chloride is considerably hydro- 
lyzed on account of the weakness of its base, and so mixing 
them gives the ferric ions an opportunity to combine with 
the acetate ions. We have before us, then, such an 
electrolytic combination of two w r eak ions as was described 
on p. 89. The H‘ ions, that in the. case of the chloride 
bring the hydrolysis to a standstill, are checked in their 
formation by the acetate ions present, and in consequence 
hydrolysis sets in to a considerably greater extent and 
leads to the formation of a much larger amount of basic 
ferric acetate along 'with undissociated acetic acid. The 
raising of the temperature favors still further this hydrol- 
ysis, so that w r e are in position to increase the concen- 
tration of the undissociated ferric hydroxide to such a 
point that the solvent capacity of the wrater for this 
substance is passed, the solution becomes saturated, and 
all further formed hydroxide must precipitate. Chromic 
hydroxide and aluminium hydroxide, as is well known, are 
also precipitated by the hydrolysis of their solutions 
containing acetate. 



144 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

To return once more briefly to heats of dissociation, the 
following deserves mention. In a number of cases it has 
been found that organic substances which are capable of 
acting either as an acid or a base must, before splitting up 
into ions, undergo a molecular rearrangement, that is, 
change into an isomeric form. We are indebted to the 
interesting investigations of Hantzsch for an entire series 
of examples of such so-called pseudo acids and bases, 
which in dissociating suffer this sort of rearrangement, and 
it seems as though in all of these electrolytic substances 
the heats of dissociation are especially large. These 
heats of dissociation manifest themselves either in the 
great variation of the degree of dissociation and the 
conductivity with the temperature, or in the heats of 
neutralization of these substances deviating considerably 
from the value 13700 cal. While there seem to be as 
yet no investigations pertaining to the latter fact, violuric 
acid and oximido-oxazolon offer two cases of the first kind, 
for which Guinchard 1 determined the variation of the 
dissociation constants with the temperature, and by 
means of this the heat of dissociation of violuric acid is 
calculated to be 3700 cal. 2 

The reaction of the intermolecular rearrangement 
preceding the dissociation must of course produce a 
certain heat effect, which appears as a part of the heat of 
dissociation. It is likely, therefore, that we are permitted 
to generalize to the extent of saying that intermolecular 
reaction and high heat of dissociation are bound together 


1 Ber. d. deutsch. chein. Ges., 32 , 1723 (1899). 

? Ibid., 33 , 393 (1900}. 



INFLUENCE OF PRESSURE AND TEMPERATURE. 145 


by a common cause. This holds without doubt not only 
for the two cases mentioned above, but in all probability 
for two other cases of high heat of dissociation given in the 
table (p. 140}, namely, hydrofluoric acid and water, since 
hydrofluoric acid is essentially present in the form of the 
molecules H 2 F 2 , not only in the gaseous state but also in 
solution, 1 and must pass through the intermolecular 
reaction 

H2F0-2HF 


in order to dissociate into the ions H' and F'. Also 
in the case of water a similar intermolecular reaction is 
more than probable, for the most varied facts have led 
to the conclusion that the water molecules in the liquid 
state are exceedingly strongly polymerized, so that they 
must likewise first break up into simple molecules of the 
formula H 2 0 in order to form H* and OH' ions. We can- 
not, however, consider this relation between high heat of 
dissociation and inner reaction as altogether general, 
because the phenols and a number of other substances 
likewise show r high heats of dissociation 2 without indi- 
cations of the probability of intermolecular reactions. It 
seems of importance, nevertheless, that the fact of the 
presence of great heat of dissociation has recently led 
Hantzsch 3 to make the interesting discovery that the salts 
of phosphorous acid are capable of existing in the form 


1 See Jaeger, Zeitschr. anorg. Chem., 27 , 28 (1901); Abeggand Herz, 
ibid.. 35 , 129 (1903). 

2 See Hantzsch, Ber. d. deutsch. chem. Ges., 32 , 3073 (1899;. 

s Zeitschr. f. Elektrochem., $, 4S4 (1902). 



146 THE THEORY OF ELECTROLYTIC DISSOCIATION.. 

of two structural isomers. That is, phosphorous acid 
also exists in two isomers of tautomeric form, and so for 
this acid it likewise appears that high heat of dissociation 
(great temperature coefficient of the conductivity, Arrhe- 
nius) is bound up with the possibility of intermolecular 
reaction. 



NON-AQUEOUS SOLUTIONS. 


All our previous discussions concerning dissociation 
have been confined to solutions in which -water was the 
solvent. This was done not alone for the reason that the 
investigation of dissociation and conductivity was first 
carried through on aqueous solutions, but because here 
we have arrived at comparatively simple results, and 
because the property of behaving as an electrolyte is an 
especially conspicuous peculiarity of the particular 
compounds when dissolved in water. The capacity for 
ionic decomposition of those substances which have been 
recognized as electrolytes in water also makes its ap- 
pearance more or less distinctly in other solvents. In 
fact, the nature of the solvent plays an exceedingly 
important role, so that it has been possible to arrange the 
various media in a series according to their “ dissociat- 
ing ” power, which series agrees for most solutes. 

One important advance in the question as to what other 
physical or chemical properties of substances the dis- 
sociating power of solvents is associated with was made 
by Nemst 1 and Thomson, and is based on a consideration 
of the dielectric constant. This constant of a medium is 
characteristic of the force with w T hich two electrically 
charged bodies within this medium attract or repel each 
other; the greater the constant is the smaller becomes the 

1 Zeitschr. physik. Chem., 13 , 531 (1894). 

*47 



14& THE THEORY OF ELECTROLYTIC DISSOCIATION, 

mutual force manifest between the electrically charged 
particles, the distance between the same remaining equal. 
Now the ions are also to be considered as such electrically 
charged parts, and accordingly the forces of attraction 
between the ions must become smaller, and their separa- 
tion from one another easier, the higher the dielectric 
constant of the medium is. It follows that the dissociation 
of substances should be especially great in that medium 
with the highest dielectric constant. In the table are 
given the dielectric constants of a number of substances 
at ordinary temperature. 

Dielectric Constants . 1 


Hydrocyanic acid, HCN 95 .0 

Hydrogen peroxide, H 2 0 2 92.8 

Water, BUO 81.0 

Formic acid, HCOOH 57.0 

Acetonitrile, CH 3 CN 36.4 

Nitrobenzene, C 6 H 5 N0 2 34 .0 

Methyl alcohol, CH3OH 32.5 

Propionitrile, C 2 H 5 CN .....* 26.5 

Benzonitrile, C 6 H 5 CN 26.0 

Ethyl alcohol, C 2 H 5 OH 22.0 

Liquefied ammonia, NHg 22.0 (—34°) 

Acetone, (CH 3 ) 2 CO 20.7 

Glycerine, C 3 H 5 (OH) 3 16.5 

Liquefied sulphur dioxide, S0 2 14.8 

Pyridine, C 5 H 5 N 12.4 

Aniline, C 6 H 5 NH2 7.2 

Acetic acid, CH 3 COOH 6.5 

Chloroform, CHC1 3 5.0 

Ether, (CoH 5 ) 2 0 4.4 

Benzene, C fi H 6 . . 2.3 


1 Schlundt, Journ. Physic. Chem., o, 165 (1901). — Drude, Zeitschr. 
physik. Chem., 23, 308 (1897). — Linde, Wied. Ann., 56, 563c 1895). — - 
Goodwin and de Kay Thompson, Phys. Review, 8, 38 (1899). — Calvert, 
Drud. Ann., 1, 483 (1900). — Coolidge, Wied. Ann., 69, 125 (1899). — 
Mathews, Bibliography of Dielect. Consts., Joum. Physic. Chem., 9 , 
667 (1905). 



NON-AOUEOUS SOLUTIONS. 149 

We see hereby that water must possess an abnormally 
high power of dissociation, since of all common solvents it 
has by far the greatest dielectric constant, and in general 
it seems that the order given in the table agrees approx- 
imately with that found in studying the ionic dissociation 
of electrolytes in various solvents by means of the con- 
ductivities. Of especial interest in this connection is an 
investigation of Centnerszwer, 1 who found the equivalent 
conductivities of potassium iodide and trimethyl sulphine 
iodide in HCN at o° to be about four times as great as 
the corresponding figures for aqueous solutions. This is 
about as great as the equivalent conductivities of the 
best-conducting electrolytes, the acids in water solutions 
at 25 0 , and even if it does not necessarily follow (see later 
on) that the dissociation is greater than in w T ater, never- 
theless this is at least possible. If such were the case, 
it would be in best agreement with what one must expect 
according to the high dielectric constant of HCN. 

On the other hand, however, these investigations have 
evidently shown that the dielectric constant cannot alone 
be the determining factor for the dissociation; moreover, 
purely chemical questions seem to be very vitally con- 
cerned. For example, according to the table of dielectric 
constants, it was to be expected that the solutions in 
benzonitrile and propionitrile would be equally dissociated 
and consequently also would show about the same 
conductivity. A comparison by Schlundt 2 of the con- 
ductivities of silver nitrate in these two dielectrically 


1 Zeitschr. physik. Chem., 39 , 217 (1902). 

2 Joum. Physic. Chem., 5 , 168 (1901). 


*50 THE THEORY OF ELECTROLYTIC DISSOCIATION. 


equal solvents, as measured by Lincoln, and Dutoit, gave 
large differences in the sense that the conductivity in 
propionitrile corresponds to a markedly greater ionic 
concentration. This fact and others are strikingly in 
accord with another assumption as to the reason for the 
difference in dissociating power, brought forward by 
Dutoit , 1 namely, the ability of the solvent to associate, 
forming polymerized molecules. Bruhl , 2 for his part, 
finds a connection with the question whether or not the 
solutions contain atoms whose valences in the compound 
are not as yet completely saturated. These two views 
seem to me to mean practically the same thing, for the 
reason that evidently the ability to associate is conditioned 
upon the presence of unsaturated valences. For the two 
nitriles mentioned, the view that polymerization plays a 
part agrees excellently with the facts, in so far as the 
researches of Ramsay and Shields 3 have shown that 
propionitrile is considerably polymerized, while benzoni- 
trile does not associate. The conclusions pertaining to 
ionic dissociation, derived from the conductivities in these 
non-aqueous solvents, are subject to considerable un- 
certainty, because in only a few' cases and for only a few 
electrolytes have wre been able to determine the values for 
Ao, the limit value of the equivalent conductivity for great 
dilution. Therefore the absolute values of the equivalent 
conductivity give only a very uncertain approximation for 
the degree of dissociation, and, strictly speaking, express 


1 Compt. rend., 125 , 240; Bull. soc. chim. (3), 19 , 321 (1898). 

2 Ber. d. deutsch. chem. Ges., 28 , 2866 (1S95); Zeitschr. physlk. 
Chem., 27 , 319 (1898); 30 , 1 (1899). 

3 Zeitschr. physik. Chem., 12 , 433 (1893). 



N6N-AQUE0US SOLUTIONS. 


151 

nothing more than that there are ions present in a solution, 
and in what way the number of ions varies in the same 
solvent with varying concentration, A comparison of dif- 
ferent solvents with one another is altogether impossible 
under these conditions, since the conductivities (see p. 29) 
are dependent not only on the degree of dissociation but also 
very much upon the mobility of the ions, and this mobility 
stands in an entirely unknown relation to the nature of the 
solvent. For that reason the attempt has been made to 
employ the other method which in the case of aqueous 
solutions has been found to be serviceable, namely, to 
measure the degree of dissociation in these solvents by 
means of the osmotic pressures (from freezing-point, 
boiling-point, or vapor-pressure determinations). And 
indeed in many cases, for example solutions in the various 
alcohols, it has been demonstrated that electrolytes, or 
substances which possess conductivity in these solvents, 
also show an increased molar number (an abnormality 
factor i> 1), as, is to be expected from the expression 
previously derived (see p. 9), 

i=i + (n— i)*a. 

However, not only have i values been found (by osmotic 
methods) which appear much too small to be made to 
harmonize with the uncertain degrees of dissociation 
derived from the a values (electrically determined), but 
in a number of cases i values have been obtained that are 
even smaller than 1, in spite of the fact that the presence 
of conductivity proves the presence of an appreciable 
degree of dissociation. These facts, which have unneces- 
sarily aw T akened doubt as to the foundations of the 



* 5 * THE THEORY OF ELECTROLYTIC DISSOCIATION. 


dissociation theory , 1 give us a clue to wherein consists the 
explanation of the anomalous behavior of the non-aqueous 
solutions. We may, for instance, maintain the relation 
between i and a if we admit the possibility that the 
undissociated molecules of the electrolyte associate to 
form polymerized molecules. Measurements have shown 
polymerization to be true of many other substances, and, 
for example, with organic acids this phenomenon has been 
repeatedly verified by the extensive investigations of 
Beckmann and others. With such association it is of 
course very* possible that the reduction in the number of 
molecules caused thereby is greater than the increase 
which has its origin in the ionic decomposition. In this 
way for the present we can give an entirely satisfactory 
qualitative explanation for the seemingly very complicated 
relations of non-aqueous solutions. Likewise in keeping 
with this, as Nemst has already stated, is the fact that the 
serial order of the dissociating powers of solvents is the 
same, whether we base it on the electrolytic dissociation 
of simple molecules into ions or the non-electrolytic 


1 See, for example, Joum. Physic. Chem., 5 , 339 (1901). The author 
Kahlenberg is an enthusiastic opponent of the ionic theory, collecting 
facts with energy and great experimental skill which appear unexplain- 
able by the theory. The possibility of suitably expanding the theory 
he unfortunately does not consider. It seems without purpose, however* 
to discard a theory so broadly established without putting a better one 
in its place. We are not accustomed to tearing down a habitable house 
because a few rooms are illy lighted, and putting ourselves out on the 
street, unless it be we can move into a better one. Any theory which 
desires to depose the one of Arrhenius has at the very outset the dif- 
ficult task of bringing into a common field of view all the manifold facts 
which Arrhenius has taught us to summarize. 



NON-AQUEOUS SOLUTIONS . 153 

dissociation of polymerized molecules into simple ones. 
Accordingly, with decreasing power of dissociation, we 
must have, in non-aqueous solvents, as compared with 
water, not only a decrease of the dissociation into ions but 
also a decrease of the dissociation of polymerized molecules 
into simple ones, or, conversely expressed, the association 
into polymerized molecules must be favored. Whether 
the in part unsatisfactory agreement between the degrees 
of dissociation of aqueous solutions, as determined by 
osmotic and electric methods, can find an analogous 
explanation is still an open question. At any rate, it is 
worthy of note, as the approximate agreement proves, 
that in water association evidently seems to play quite a 
secondary role. 

It is well to add that the probability of a participation 
of the solvent in the process of ionic dissociation, such as 
the formation of an addition product of solvent and ions, 
is variously supported by the experience with non-aqueous 
solutions. The assumption that the presence of free 
valences assists in determining the dissociation led Briihl to 
the conclusion that the ions are hydrated; and the results 
of Walden, that iodides and sulphocyanates have especially 
high solubilities in SO2, with peculiar colorations of their 
SO2 solutions, indicating the formation of new com- 
pounds , 1 point to the same conclusion. That this agrees 
with the facts is shown in a research by Fox , 2 who also 
proved the existence in aqueous solutions of such complex 
ions of the halogens and other anions with S 0 2 . The 


1 Zeitschr. physik. Chem., 42 , 432 (1903). 

1 Dissert., Breslau, 1902 ; Zeitschr. physik. Chem., 41 , 458 (1902). 



154 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

general result of the knowledge gathered thus far with 
non-aqueous solutions may for the present be summed up 
to the effect that the simplicity of conditions prevalent in 
aqueous solutions is here very much complicated by the 
phenomenon of the association of the non-ionized portion 
of the electrolyte. Hence the problems to be solved in 
this connection are, first, the determination of the amount 
of association of the electrolytes and, next, the Kohlrausch 
law for the additive nature of Jo- According to the 
equilibrium laws, it is no doubt permissible to assume 
that in non-aqueous solutions the ionic splitting-up, 
though in any case probably very slight, is caused by the 
great reduction, due to association, of the active masses 
of the simple molecules that alone are capable of ionizing. 
No doubt we might assert with equal right that in water 
the active mass of the undissociated molecules becomes too 
small, on account of the great ionization, to allow an 
appreciable association, since this latter is again pro- 
portional to the active mass of the simple molecules. 
But in many instances, for example with the organic 
carboxylic acids, the electrolytic dissociation is so slight 
even in water that the active mass of the undissociated 
molecules is not appreciably affected by it; that is, even 
if ionization is wanting, association would not increase in 
a sensible degree. ' So that in such cases we also find the 
association in water exceedingly small as compared with 
the other solvents, for which, according to the researches of 
Beckmann, the organic acids and alcohols already alluded 
to give the best evidence. This leads one to the assump- 
tion that the phenomenon of association is the chief 
phenomenon, and that ionization in non-aqueous solu- 



NON-AQUEOUS SOLUTIONS. 155 

tions is so small, as a rule, essentially because of the great 
amount of association. 

Finally, we may consider that case which deals with the 
amount of dissociation in pure substances as one of non- 
aqueous solvents, or, as we may term it, one of self -disso- 
ciation. Here it is a general rule that the electrolytic 
conductivity of pure substances is without exception 
exceedingly small, though, according to Walden, 1 it seems 
to increase with the polar difference of the atomic com- 
ponents. The amount of self-dissociation of water and 
the conductivity resulting therefrom have already been 
discussed (see p. 56). Conductivities of about the same 
order of magnitude have been observed for a whole series 
of other liquids of greatest possible purity, such as methyl 
and ethyl alcohol, sulphur dioxide, liquid ammonia, etc. 
But judging by the experiences of Kohlrausch and 
Heydweiller with water, it is difficult to say in how far 
the conductivities thus obtained are those of the absolutely 
pure substances, or to what extent possibly they are 
conditioned upon electrolytic impurities. It is worth 
mentioning in this connection that such substances as 
sulphuric acid, hydrochloric acid, etc., which in aqueous 
solutions belong to the best electrolytes, in the pure liquid 
form, on the contrary, possess an exceedingly small con- 
ductivity. The only known exceptions to this are the 
salts which in the melted state, that is as liquids, are very 
good electric conductors. It may be that this behavior 
is connected with a large dielectric constant of the salts in 


1 Zeitschr. anorg. Chem., 25 , 225 (1S90); see also Abegg, Christiania 
Vidensk. Selsk. Skrifter, 1902, No. 12, p. 8. 




I5 6 the theory of electrolytic dissociation . 


the molten state, although it has not been possible thus far 
to determine the same. In the solid state they show 
dielectric constants between 6 and 7, while all other solid 
substances, even water in the form of ice, 1 show only 
1 to 3; and since liquids and melts always have higher 
dielectric constants, it does not seem improbable that 
the salt melts are exceedingly strong dielectrics. From 
this follows the probably very great self-dissociation which 
lies at the basis of the good conductivity of the salts, and 
according to the Thomson-Nemst rule is determined by 
their high dielectric constant. It is true at the high 
temperatures of the melted salts the ionic mobilities may 
be so great that it may not be necessary to consider the 
good conductivity as due to a marked dissociation. 
Especially in the case of water, where, as a result of the 
high dielectric constant, one might expect a greater self- 
dissociation, the very great association is in all likelihood 
essentially to blame, which takes away the simple mole- 
cules, the real material for the ionization. 

The determination of electrolytic dissociation based 
solely on conductivities, however, without knowledge of 
the specific resistances which the medium opposes to the 
transport of ions, must always be looked upon as extremely 
uncertain. So, for instance, the view has been frequently 
expressed or entertained that the salts in the solid form are 
not electrolytes, or at most possess only a very slight 
electrolytic dissociation. This is probably based on the 
observation that the solid salts have an exceedingly slight, 


1 See Abegg, Wied. Ann., 65 , 229 (189S), and Zeitschr. f, Elektro- 
chem., 5 , 353 (1899). 



NON-AOUEOUS SOLUTIONS . 


I 57 


but nevertheless perfectly definite, conductivity, as 
determined by Warburg. The color of many solid salts, 
especially when they crystallize as hydrates, is very often 
identical with the specific ionic color of the chromophore 
constituent, and thus makes the presence of ions probable. 
The small conductivity in spite of the ions may easily be 
explained by the enormous frictional resistance to which 
the moving particles in their solids are subjected. The 
greatest promise of success in penetrating into the quanti- 
tative dissociation relations of non- aqueous solutions 
seems to be offered by studies in gradually varying 
solvents, as an example of which the research of Wolf, 
mentioned above (p. 132), may be cited. The first in- 
vestigation with this object in view comes from Arrhe- 
nius, 1 who studied the influence of small amounts of 
non-conductors, such as alcohol, sugar, etc., on the 
conductivity and dissociation of various electrolytes. 
The chief result of this investigation is the establishing of 
the fact that the changing of the solvent also produces 
changes in the degree of dissociation, but of very different 
amounts, according to the nature of the electrolyte. The 
strongest dissociated electrolytes are affected very little 
in their degree of dissociation by slight changes of the 
medium, while the weak electrolytes are extremely 
sensitive to such changes, their degree of dissociation 
being reduced. Cohen 2 has confirmed the results of 
Arrhenius for strong electrolytes in the entire interval of 
solvent produced by mixing alcohol and water in various 


1 Zeitschr. physik. Chem., 9 , 487 (1892). 

2 Ibid., 25 , 1 (1S99). 



I 5 8 THE theory of electrolytic dissociation. 

proportions. He found that the degrees of dissociation 
in these cases do not seem to be influenced by the medium, 
at least in so far as the conductivity represents a correct 
measure of dissociation. We see, therefore, that it is less 
the nature of the solvent than that of the dissolved sub- 
stance which is of influence here, and this leads us to a 
final consideration, whose subject is the important 
question, What sort of regularities exist between the 
endeavor of the electrolytes to dissociate and the nature 
of their components ? 



CHEMICAL NATURE AND IONIZATION 
TENDENCY OF THE ELEMENTS. 


It has become apparent that of the elementary substances 
an entire series appears altogether, or at least by evident 
preference, in the form of ions. Thus, for example, there 
is not a single compound of the alkali and alkaline-earth 
metals which does not contain these metals for the greater 
part as independent ions, while others, again, such as most 
of the elements belonging to the carbon and nitrogen groups 
of the periodic system, are as good as unknown in the 
form of elementary ions. Furthermore, the difference in 
capacity for forming positive and negative ions is very 
striking: elements such as fluorine and chlorine never 
appear as positive ions, while the elements of the first two 
groups of the periodic system act exclusively as positive 
ions In the middle groups the tendency to form ions 
disappears more and more, in place of which these 
elements assume an amphoteric character in that they 
give evidence of a participation in the ion formation of 
others combined with them, even though it has not been 
possible thus far to confirm an independent formation of 
ions. 

A good illustration of the preceding statement is offered 
by nitrogen, which combined with four H atoms furnishes 

159 



i6o the theory of electrolytic dissociation . 


the cathion NH 4 ‘ (ammonium), while in the form of HN 3 
orHN0 3 it produces the anions N 3 ' andN0 3 ' respectively. 
Likewise phosphorus forms cathions in the phosphonium 
compounds as well as anions in the acids of phosphorus 
and the phosphides, 1 and again, sulphur shows a varying 
polar behavior in the sulphine bases and in sulphides or 
sulphuric acids. Furthermore, we are familiar with iodine 
independent, and also as an anion-former in many com- 
plex combinations, but nevertheless in addition it is' capa- 
ble, as taught by the existence of iodonium bases, of enter- 
ing into the garb of a cathion. 2 It is possible in general to 


1 Schenck, Ber. d. deutsch. chem. Ges., 36 , 979 (1903). 

2 Of such amphoteric electrolytes there are quite a number; thus, 
almost all hydroxides of the weak positive elements along with oxygen 
can form anions, so that their hydroxides are capable of producing 
simultaneously cathions of the element and anions of its oxygen complex. 
Among others, this is known to be the case with Pb(OH) 2 , Al(OH) 3 , 
Cr(OH) 3 , As(OH) 3 , As(OH) 5 , Be(OH) 2 , Sn(OH) 2 , Ge(OH) 2 (cf. 
Hantzsch, Zeitschr. anorg. Chem., 30 , 289 [1902]; McCay, Journ. Amer. 
Chem. Soc., 24 , 667 [1902]), and is deserving of interest because upon 
dissociation OH' ions are formed on the one hand and H* ions on the 
other, whose mutual concentrations in the presence of water are limited 
by the water constant k w (see p. 55). In the coupled equilibrium, 
such as, for instance, 

Pb** + 2OH' ^ Pb(OH) 2 H* -f Pb 0 2 H', 

an addition of strong bases (OH' ions) retards the production of Pb** on 
account of the common OH' ions, that is, forces back the basic function 
of the hydroxide, while the consumption of H* ions (as a result of the 
formation of water) increases the concentration of Pb 0 2 ". In short, 
the add nature (anion formation) is enticed forth, so to speak, which 
agrees with the experience on other amphoteric substances. It is also 
worth while to call spedal attention to the reaction, mentioned above 
(see p. 1 17), of boric acid with fluorides, in which the OH' concentration 
for the complex fluorides is brought forth by the consumption of the 



CHEMICAL NATURE OF THE ELEMENTS . 161 

confirm, by reference to the periodic system, the fact that 
the tendency in each principal group to form cathions 
increases with increase in the atomic weight, or, what is 
the same thing, the tendency to form anions increases 
with diminishing atomic weight. In the horizontal rows, 
however, the tendency toward the formation of cathions 
diminishes with increasing atomic weight. Quantitative 
knowledge, at least of an approximate nature, is given 
by the study of the electromotive activity of the elements 
and of the decomposition voltages, or the amount of 
electrical energy necessary to deprive an ion of its charge. 
Not wishing to go farther here into these interesting 
questions, suffice it to refer again to the previously men- 
tioned articles of Abegg and Bodlander on Electro- 
affinity, and Abegg on Valence. 1 Be it only noted that 
the tendency to form ions, or the electro-affinity, manifests 
itself also in the solubility relations 2 of the electrolytes 
and their capacity for forming complexes; that is, the 
disinclination to form ions is often associated with slight 
aqueous solubility or with great tendency to enter into 
complex ions, whereby the compound avoids the necessity 
of being subjected to the dissociating force of the water. 

The observation that all reactions in which ions 
participate in measurable amounts — even the hydrolytic 


borate ions. This indicates the presence of a weak basic nature in 
boric acid. A number of organic amphoteric electrolytes have been 
more carefully studied by Winkelblech (see p. 55, alanine, amido. 
benzoic acid). 

1 Zeitschr anorg Chem., 20 , 453 (1899); Christiania Vidensk. Selsk. 
Skrifter, 1902, No. 12, p 8. 

* Cl. Imnjerwahr, Zeitschr. f. EJektrochem,, 7 477 (1901), 



162 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

actions of the exceedingly weakly dissociated water — 
proceed to their equilibrium with an immeasurably 
great velocity has induced the assumption that indeed 
every capacity to react is to be attributed to the presence 
of ions. A basis for this assumption has been thought to 
exist in the fact that reactions between non-electrolytes 
usually proceed with extreme slowness, corresponding to an 
immeasurably small, but not altogether lacking, dissocia- 
tion. We may, by way of illustration, consider the hydrol- 
ysis of stannous chloride and stannic chloride. Both are 
dependent upon the action of the OH' ions of the water 
on the tin of the compound. The SnCl 2 , which shows 
the presence of appreciable quantities of tin ions by the 
fact that during electrolysis metal separates as a result of 
the discharge of these ions, attains its state of hydrolytic 
equilibrium momentarily; while, on the other hand, the 
SnCU contains no demonstrable — or rather extremely 
small — amounts of tin ions, and accordingly its hydrolysis 
goes on very slowly, as traced by Kowalevsky . 1 Analogous 
facts hold for the hydrolysis of PtCU and AUCI 3 , also for 
the hydrolytic splitting-up of the esters , 2 as observed by" 
Kohlrausch . 3 The exceedingly interesting investigations 
of Brereton Baker 4 on the failure of gases to react when 
absolutely dry, as well as the non-dissociation of NH 4 C1 
and of Hg 2 Cl 2 , the failure of NH 3 and HC1, H 2 and Cl 2 , 


1 Zeitschr. anorg. Chem., 23 , i (1900). 

2 Zeitschr. physik, Chem., 36 , 641 (1901). 

3 Ibid., 33 , 1257 (1900). 

4 Joum. Chem. Soc., 73 , 422; 77 , 646; 81 , 400 (1899-1902). See %lso 
Noyes, Zeitschr. physik. Chem., 41 , n (1902). 



CHEMICAL NATURE OF THE ELEMENTS . 163 

and H 2 and 0 2 to react, and lastly the lack of electric 
conductivity, which we attribute to ionization— all this 
speaks likewise in favor of the assumption that capacity 
for reaction is due to ions. 

A more recent investigation of Kahlenberg, 1 which offers 
as evidence against the above the instantaneous pre- 
cipitations in non-aqueous solutions possessing no demon- 
strable conductivity, should be completed in the direction 
of striving to attain by all known means the absolute 
dryness so difficult to accomplish, as the experience of 
Baker shows. Until that has been done, we may only 
conclude that very great reaction velocities can be reached 
even with quantities of ions so small as to be beyond 
detection, and it would depend upon penetrating into the 
region of these small ionic concentrations, thus far in- 
accessible, in such a manner as to measure their small 
concentrations. 

The dissociation theory has taught us to consider from 
a common point of view and to understand in their mutual 
relations an immense number of facts coming from the 
apparently most diverse regions of chemistry. Doubtless 
an equally great number of problems this theory has 
presented to science and has helped or helps in their 
solution. Its successes in the field of chemistry are no 
greater than in the field of physics, where the brilliant 
researches of Nemst have solved with its assistance the 
theory of diffusion and the hundred-year-old problem 
of the voltaic chain. Yes, one cannot resist the impres- 
sion that the future of the theory will lead us directly to 


1 Joum. PKysic. Chem., 6, 1 (1902). 



164 THE THEORY OF ELECTROLYTIC DISSOCIATION. 

the ultimate questions of chemistry, the essence of 
valence and the affinity forces; and so one can maintain 
that this conception of Arrhenius is one of the most 
significant and fruitful with which theoretical chemistry 
has ever been favored. 



INDEX. 


PAGE 

Abegg. 123 

degrees of dissociation from freezing-points and equivalent con- 
ductivities 45 

dilution equation, concentrated solutions \ 132 

see Bersch - 116 

valence 161 

and Boildnder 127 

electroaffinity 161 

Absorption spectra of ions, Ostwald 12 

Acid reaction on mixing neutral and alkaline compounds 117 

Acids, and bases, neutralization of 58 

avidity of, Ostwald , Arrhenius , Wolf 99 

catalysis of cane-sugar in the presence of neutral salts 72 

conductivity temperature coefficients 36 

definition of 5 

dibasic dissociation 51, 53 

influence of substitution on strength of 47 

liberation from salts 75 

pseudo, Hantzsch 144 

Additive properties of solutions 12 

Alkaline reaction of difficultly soluble oxides and neutral salts, Bersch , 

Ahegg, 1 16 

Ammonia, solvent action of 115 

Amphoteric electrolytes 160 

Anion, slowest 35 

Anions, mobilities of organic 34 

Anomaly of strong electrolytes 121 

*65 



INDEX. 


1 66 

PAGE 

Arrhenius 1,2, 8, 9, 16, 25, 66, 141, 164 

avidity 98 

conductivity and dissociation in mixed solvents 157 

constants of inversion 73 

degrees of hydrolysis 91 

electrolytic dissociation 3,4 

equilibrium relations among electrolytes 69 

heats of dissociation 140 

temperature equation of 'dissociation constant 139 

theory of isohydric solutions 62 

and Fartjung , influence of pressure on dissociation 134 

Walker, hydrolysis of salts of two weak ions 91,92 

Association and ionization 154 

Avidity 95 

theory of 96 

Baker, non-reactivity of dry substances. 162 

Bases, and acids, neutralization of 58 

definition of 5 

influence of substituents on strength of 51 

liberation from salts. 75 

Basic salts, formation of, by hydrolysis up 

Bassett , see Dorman 126 

Baur , see Arrhenius 140 

Beckmann 154 

polymerization in solution 152 

Bersch, Abegg , alkaline reaction by interaction of neutral compounds 117 

BUtz, dissociation of caesium nitrate 127 

and Meyer , conductivity of caesium nitrate 128 

Bodlander 120 

see Abegg. 127,161 

Boiling-point, see RaouU . 

Bredig , action of substituents on strength of bases 51 

confirmations of the dilution law 45 

hydrolysis p X 

hydrolytic decomposition and ionic mobilities 86 

inner complexes 125 

mobilities of inorganic and organic ions 35 

see Oshvald. ** 



INDEX. 


167 

PAGE 

Bruhl, hydrated ions 153 

unsaturated compounds and ionization 150 

Buckingham, fluorescence of ions 13 

Buff 3 

Calculation of, absolute mobility * 31 

electric mobility 32 

equivalent conductivity at infinite dilution 34 

maximum equivalent conductivity 32, 34 

Cannizzaro , Kopp , and Keknle , abnormal vapor densities 7 

Carbonates, solution of 115 

Catalysis of cane-sugar by acids in the presence of neutral salts. 73 

Cathion, slowest 35 

Centner szwer, equivalent conductivities in HCN 149 

Chemical nature and ionization tendency of the elements 159 

Chlorine titration, Mohr's method 111 

Volhard's method no 

Clausius 3 

Cohen , mixed solvents 157 

Color of ionic solutions 12, 13 

Complexes, formation of, Hittorj 125 

inner 125 

solubility, and electroaffinity 161 

with S 0 2 in H a O solutions, Fox 153 

Concentration chains, degree of ionization by means of 28 

John 126 

Concentration, changes in the vicinity of the electrode, Hittorj 29 

high, and conductivity 132 

in equivalents 20 

in normals 20 

' relation to conductivity. 43 

sign 1 

Concentrated solutions, dissociation of, Rudorj , Wolf 132 

Concentrations, isohydric 64 

Conductivity, and dilution constant 42 

and salts, Hittorj 4 

at infinite dilution 20, 22 

equivalent 20 

in benzonitrile and propionitrile, Schlundt, Lincoln, Dutoii. 


149 



i6S 


INDEX. 


page 

Conductivity, in HCN, and dielectric constant, Centner sz-mr 149 

in mixed solvents, Arrhenius 157 

maximum equivalent, calculation of 32, 34 

of isobydric solutions, Wakeman 66 

of pnre fused salts 155 

of pure water, Kohlrausch and Heyd'iveiller 56 

of solid salts 157 

specific 19 

temperature and mobility 36 

temperature coefficient of water 142 

variation with concentration 43 

Constant, hydrolytic 80 

hydrolytic, of KCN, Shields 82 

of dissociation, see Dissociation constant. 

of inversion of adds in the presence of neutral salts . . 73 

dielectric 148 

Criticisms of ionic theory 14, 15 

Current transport, mechanism of, within the solution 30 

Davy 3 

Decomposition voltages . 161 

Degree of, dissociation from freezing-point and equivalent con 

ductivity, Abegg 45 

hydrolysis , ; , . 80 

Shields , Walker gj- 

ionization 24 

by concentration chains, Jahn 28 

Denison , see Steele ^4 

Dibasic acids, dissociation of . 5D53 

Dielectric constant, and dissociating power of solvents 147 

of salts 4 

Diffusion, chains, Nernst. j5 

of electrolytes, Nernst . . 

Dilution law, Ostivald, Starch . , van’t Hoff 39, 41, 130 

and caesium nitrate .. t I2 g 

concentrated solutions, Abegg , Rudorf. 3:32 

confirmation of ^ 

exceptions I2I 

theoretical basis, Roloff 



INDEX 


169 


PAGE 

Dissociating power of solvents 147 

Dissociation, and i factor. 25 

and temperature 135 

and solvent 151 

and medium 133 

based on conductivity 156 

in mixed solvents, Arrhenius , Cohen , Wolj 157 

isotherm 131 

non-electrolytic 152 

of caesium nitrate according to conductivity and freezing-point. . . 128 

of concentrated solutions, Rudorf , Wolf 132 

of dibasic acids, Ostwald 51, 53 

of mixed electrolytes 67, 69 

of pure water. 56 

of solute and association of solvent, Briihl , Dutoit 150 

of water, variation with temperature, Kohlrausch and Heydweiller 

60, 142 

pressure influence on, Fanjung 134 

Dissociation constant 40, 46 

and chemical nature 47 

and hydrolytic constant 82 

degree of dissociation, equivalent conductivity 41, 44 

influence upon, of substitution 47 

of dibasic acids 53 

of HCN, Walker 82 

of RbN 0 3 127 

physical significance of, Ostwald 46 

strong electrolytes, Jahn , Kohlrausch , Rudolphi , van’t Hoff 

123, 129 

temperature equation of, Arrhenius 139 

Dissociation constants 47,48 

Dissociation degree 24 

and dilution constant 41 

by means of isohydric solutions 69 

from freezing-point and equivalent conductivity, Abegg 45 

Dissociation heat 136, 140 

and pseudo acids 144 

variability with temperature, Kohlrausch and Heydweiller 140 

Donnan, Bassett , and Fox, inner complexes 126 



110 INDEX. 

PAGE 

Diitoit, association of solvent and dissociation of solute 150 

see Schliindt 149 

Electroafimity, Abegg and Bodlander 161 

solubility and complexes 161 

Electrode, concentration change in the vicinity of 29 

Electrolytes 2, 3, 4 

amphoteric 160 

anomaly of strong 121 

dissociation constant of- strong, Jahn 123 

dissociation of mixtures 69 

equilibria among several 61 

extremely weak 55 

strong. 26 

strong, equation for dissociation constant of, KohlrauscJi , Rudd phi, 

van't Hoff. ’ 129 

strong, obeying dilution law, Ostwald 124 

weak 26 

with negative temperature- coefficient of conductivity 140 

Electrolytic, equilibrium relations, Arrhenius 61, 69 

equilibria, heterogeneous 105 

mobilities, Kohlrausch 3 2 j33 

Electrolytic dissociation, see Dissociation, Arrhenius 4 

fundamental conceptions of the theory 1 

Elements, capacity for forming positive and negative ions 159 

v. Ende 120 

Energy, ionic consumption of, during electrolysis 3 

Equation, a and i 25 

a and A 24 

dissociation constant, strong electrolytes, Kohlrausch , Rudolphi , 

% vanH Hoff 129 

of Jahn - 123 

of van 1 ! Hoff 59 

Equilibria, among ions 37 

among mixtures of electrolytes of equal strengths with common ions 74 

among several electrolytes 61 

heterogeneous, and hydrolysis 119 

heterogeneous electrolytic 105 

Equilibria in the hydrolysis of KCN 78 



INDEX. 


171 

PAGE 

Equilibrium of hydration 17 

Equivalent conductivity (A) 20 

degree of dissociation, dissociation constant 44 

freezing-point, degree of dissociation 45 

maximum, calculation of 32, 34 

Euler , hydration of halogen ions 35 

Fanfung , pressure and dissociation 134 

Faraday / 3, 14 

F, unit electrochemical quantity of electricity 14, 21 

law of 14 

Fluorescence of ions 13 

Fox , complexes with S 0 2 in aqueous solutions 153 

see Denman 126 

Freezing-point, equivalent conductivity and degree of dissociation 45 
see Raoult. 

Goodwin 1 19 

testing of solubility law : 107 

Guinchard, dissociation constants of violuric acid and oximido-oxa- 

zolon, variation with temperature 144 

Halogens, hydration of, Euler 33 

Hantzsch 160 

isomers of phosphorous acid 146 

pseudo acids 144 

Heat of dissociation, and isomeric forms 145 

Arrhenius , Baur , Thomsen 140 

methods of determining 136 

Heat of neutralization, methods of determining 138 

Hess 39 

Thomsen 137 

Helmholtz 3 

Hess. 60 

heat of neutralization 59 

thermoneutrality of salts 60 

Heterogeneous electrolytic equilibria 105 

Heydweiller i see Kohlrausch. 56,60, 140, 155 



INDEX . . 


172 

PAGE 

Hiitorf 3, 13 

apportionment of the ions in transporting current 29 

determination of nature of ions 4 

electric conduction of salts . 4 

formation of inner complexes 125 

ratio of rates of migration of opposite ions 30 

van't Hoff , see under V. 

Holborn , see Kohlrausch 45 

Hydrate, equilibrium 17 

theory 16 

theory and i factor 18 

Hydrates, ionic * 126, 127 

H} r dration, and mobility of inorganic ions 35 

of the halogen ions, Eider 35 

Hydrogen ions, characteristic reactions of 5 

in organic compounds, conditions for production of 51 

Hydrogen sulphide, ionization of. Walker 112 

Hydrolysis 76 

an ionic reaction, Kovalevsky 162 

and heterogeneous equilibria up 

and precipitation 95, 143 

and water constant 142 

degree of 80 

degree of, Bredig, Shields, Walker p X 

of KCN, Shields g 2 

of salts of two weak ions 89 

outward recognition 84 

quantitative relations of, Ley p 2 

reduction of 

Hydrolytic, decomposition X 4 

relations from the side of products of hydrolysis 95 

Hydrolytic constant 80 

and dissociation constant g 2 

of oxides II4 

Hydrolytic constants. Walker 84 

Hydroxides, precipitation of, and hydrolysis !43 

solubility of II4 

Hydroxyl ions, characteristic reactions of 5 



INDEX *73 

PAGE 

i factor . 8, 9 

i factor, arid hydrate theory* 18 

and number of ions, Taylor 18 

osmotically and electrically measured 25 

Immerwahr 1 19 

Indicators 100 

selection of 104 

Influence of pressure and temperature on dissociation 134 

Inversion constant for acids in presence of neutral salts, Arrhenius. 73 

Ion, slowest anion 35 

slowest cathion 35 

Ionic, decomposition products, separation of 15 

equivalent, quantity of electricity carried by 29 

hydrates 126,127 

mobility, temperature influence, Kohlrausch 33, 35 

product 106 

reactions 11,162 

signs 1 

theory, criticisms of 14, 15 

Ionization, conception of 14 

degree of 24 

degree of, by concentration chains 28 

history of, Roloff 3 

see Dissociation- 

tendency and chemical nature 159 

tendency and solubility 161 

Ions 3 

and hydrolysis 92 

and law of mass action 39 

and periodic system 161 

color of, Ostwald 12, 13 

determination of nature of 4 

energy consumption of, during electrolysis, Buff , Clausius , 

Helmholtz *. 3 

equilibria among 37 

fluorescence of, Buckingham 13 

hydration of the halogens, Euler 35 

hydrogen 5 

hydrogen, discharged, Ostwald 16 



i74 


INDEX. 


PAGE 

Ions, hydrogen, in organic compounds 51 

hydroxyl 6 

independent nature of 14 

inorganic, mobility and hydration of 35 

law of Kohlrausch 21 

mobility of, Bredig, Kohlrausch 29, 33, 35 

mobility of organic 34 

number of, normal value of osmotic pressure, and i factor 18, 19 

positive and negative, from elements 159 

reactions, and reaction velocity 162 

transference numbers of 31 

Isohydric concentrations 64 

solutions, conductivity of 66 

solutions, theory of, Arrhenius 62 

Isomeric forms and heat of dissociation 145 

Jaeger , solubility of HgO in H 2 F 2 114 

Jahn 121, 129 

degree of ionization by concentration chains 28 

deviations from the dilution law 122 

equation for E.M.F. of concentration chains 126 

equation of 123 

see Nernst 122 

Kohlenberg 152 

reactions in non-aqueous solutions 163 

Kekule 7 

Kohlrausch 34 

electrolytic mobilities 33 

equation for the dissociation constant 129 

equivalent conductivity and dilution 20 

hydrolysis of esters, an ionic reaction 162 

influence of temperature on ionic mobility 33 7 35 

law of the independent migration of ions 21, 22 

relation of concentration to conductivity 43 

and Ueydweiller 133 

conductivity of pure water 56 

conductivity temperature coefficient of water 141 


variability of heat of dissociation and temperature, 140 



INDEX. 175 

PAGE 

Kohlrausch and Heydweiller, water dissociation and temperature.. 60 

and Holborn , confirmations of the dilution law 45 

and Rose 120 

Kopp . . 7 

Kowalevsky, hydrolysis, an ionic reaction 162 

Kuster and Tkeil 120 

Law of, Faraday 14 

Kohlrausch 21 

mass action and caesium nitrate 127 

mass action applied to ions 39 

Raoult 6 , 7 

solubility influence, Nernst 107 

Le Chalelier, principle of 134, 135 

“Lehrbuch der Allgemeinen Chemie,” Ostwald 2 

Ley f quantitative relations of hydrolysis 92 

Lincoln , see Schlundt 149 

Mass-action law and ions 39 

McCay 160 

Mechanism of the current transport within the solution 30 

Meyer , see Biltz 128 

Migration law of ions, Kohlrausch 21 

Migration of opposite ions, ratio of rates, Hittorf 30 

Mixtures of salts. 10 

Mobilities, electrolytic 33 

Mobility, absolute, calculation of 31 

conductivity, and temperature 36 

electrolytic, calculation of 4 . . . 32 

ionic, of hydrolyzed salts, Bredig 86 

of ions 29 

of ions, Bredig 35 

of ions, temperature influence on 33, 35 

of organic anions 34 

Mohr's chlorine titration method no 

Molar depression of the freezing-point of water 6 

Mole 9 

Molecular weight in solution, Raoult 2 

Molecule number methods . .7 



i7 6 


INDEX. 


PAGE 

Nernst 17, 28, 106, 152, 163 

electrolytic diffusion 16 

see Ostwald 15, r6 

solubility law 107 

and Thomson's rule 147 

and Jahn, interaction of ions and undissociated molecules 122 

Neutral compounds, alkaline reaction of, by interaction „. 116 

Neutralization, acids and bases 58 

analogy to 76 

by use of borax 88 

heat, Hess 59 

heats, Thomsen 137 

Non-aqueous solutions 147 

reactions in, Kohlenberg 163 

Non-electrolytic dissociation 152 

Noyes. . . . 120 

inner complexes 125 

solubility of thallous chloride 107 

testing of solubility law 107 

Organic anions, mobility of 34 

Osmotic pressure, methods for determining molecular weight, see 

Raonlt 2 

normal value of, and number of ions 18 

Ostwald , absorption spectra of permanganates 12 

additive properties 1 

avidity 96,98,99 

Bredig , degree of dissociation, equivalent conductivity, disso- 
ciation constant 44 

confirmations of the dilution law 45 

dilution law *. 39,41 

dissociation of dibasic acids 51, 53 

equivalent conductivity. 20 

grouping electrolytes as to strength 25 

ionization of water. . 57 

“Lehrbuch der Allgem einen Chemie” 2 

physical significance of the dissociation constant 46 

solubility product - 106 

strong electrolytes obeying dilution law .>■ 124 ‘ , 



INDEX . 


177 


PAGE 

Ostwald, “Zeitschrift ftir physikalische Chemie ,, 2 

and Nernst , discharged hydrogen ions 16 

ionic charges 15 

Oxide of mercury, solubility in H_F 2 114 

Oxides, solubility of 114 

Periodic system and ions 161 

Phosphorous acid, isomers, Hantzsch 146 

Polymerization, and ionization 150 

in solution, Beckmann 152 

Precipitates, conditions of formation 108 

conversion of 109 

solution of 1 15 

Precipitation, by hydrolysis 95 

of sulphides 112 

Pressure, influence on dissociation, Fanjung 134 

Principle of Le Chatelier 134 

Pseudo acids, Hantzsch 144 

Ramsay and Shields , polymerization of liquids 150 

Raoult - 2,7,8,17 

freezing-point, boiling-point, vapor-pressure methods 7, 151 

law of 6, 7 

molecular weight in solution 2 

Ratio of the rates of migration of the opposite ions 30 

Reaction velocity and ions 162 

Reactions, ionic 11, 162 

and ions. Baker 162 

Reicher , see va?i't Hof 25, 41 

Rolof, history of ionization 3 

theoretical basis for dilution law 131 

Rose f see Kohlrausch 120 

Rudolphi , equation for dissociation constant of strong electrolytes. . 129 

Rudorf , dissociation of concentrated solutions 132 

Rule of Nernst and Thomson 147 

Sackur 129 

‘‘Salt,” conception of 4 

Salt, mixtures : ig 



178 


INDEX. 


PAGE 


Salts, conductivity of pure 155 

conductivity of solid, Warburg 157 

conductivity temperature coefficients 36 

dielectric constant of 155 

formation of basic by hydrolysis 119 

heats of dissociation of, Arrhenius . 141 

hydrolysis of. Shields 87 

ionic mobility of hydrolyzed, Bredig 86 

of two weak ions, hydrolysis of 89 

reactions of hydrolyzed 93 

self-dissociation of 156 

thermoneutrality of 60 

“Sammlung chemischer und chemisch-technischer Vortrage” . . . . 2 

Sclilundt , Lincoln , DiUoit , conductivities in benzo- and propionit- ile 149 
Self -dissociation 155 


of salts. 


156 


Sherrill. 


120 


Shields , degrees of hydrolysis 91 

hydrolysis of salts 87 

hydrolytic constant of KCN 82 

ionization of water 57 

see Ramsay 150 

Sign, concentration 1 

Signs, ionic 1 

Solubility, electroaffinity, and complexes 161 

law, Nernst , 107 

of oxide of mercury in 114 

of oxides and hydroxides 114 

of sulphides 113 

of thallous chloride, Noyes 107 

product, Ostwald 106, 108 

products, numerical values 120 

Solution, of precipitates 115 

polymerization in, Beckmann. 152 

Solutions, additive nature of properties 12 

dissociation of concentrated, Rudorf , Wolf 132 

non-aqueous 147 

theory of isohydric, Arrhenius 62 

Solvent action of ammonia, 115 



INDEX . 


179 


PAGE 

Solvent, and dissociation 151 

and solute, union of, Briihl , Walden .. . 153 

Solvents, dissociating power of 147 

electrolytic conductivity of pure, Walden 155 

mixed 157 

Sounding instrument, chemical in 

Specific conductivity (x) 19 

Specific gravity, additive nature of 13 

Steele, inner complexes 125 

and Denison , transference numbers of ions with greater valence 

than one 34 

Storch , dilution law 130 

Strength, of acids, influence of substitution on 47 

of bases, influence of substitution on 51 

Strong electrolytes - 26 

Substituents, position of, and strength of acids 50 

Substitution, influence on dissociation constant 47, 51 

Sulphides, grouping as to solubilities 113 

precipitation of 112 

Taylor , sodium mellitate and number of ions 18 

Temperature, coefficients of conductivity of acids and salts 36 

conductivity, and mobility 36 

influence on dissociation 134 

influence on ionic mobility. 33, 35 

Theory, fundamental conceptions of electrolytic dissociation 1 

hydrate 16 

of avidity '. 96 

of solution, van’t Hojf 2 

Thermoneutrality. 136 

of salts, Hess 60 

Thiel , see Kiister : 120 

Thomsen > 60 

avidity 96 

heats of neutralization 137 

see Arrhenius 140 

Thomson, see Nerns X47 

Transference number, of the ion .■ 30 

of ions with greater valence than one, Steele and Denison 34 

Transport of current within the solution 30 



i8o 


INDEX . 


PAGE 

Valence, A begg. 161 

Valson, additive nature of specific gravity 13 

van’tHoff 2, 25,57, 131 

dilution law of 130 

dissociation constant equation 129 

equation 59 

i factor 8 

solution theory 2, 7 

and Reicher, degree of dissociation and dilution constant 41 

Vapor densities, abnormal 7 

Velocity of reaction and ions 162 

Volhard , chlorine titration in 

Voltages, decomposition 161 

von Ende . 120 

Wakeman?i y conductivity of isohydric solutions 66 

W alden 45 

electrolytic conductivity of pure solvents 155 

union of solvent and solute 153 

Walker. 87 

degrees of hydrolysis qx 

dissociation of HCN 82 

hydrolytic constants * 84 

ionization of hydrogen sulphide 112 

see Arrhenius. qi 

Warburg } conductivity of solid salts 137 

Water constant and hydrolysis 142 

Water, dissociation of, Ostwald , Shields , Wijs 57 

dissociation, variation with temperature 60 

electrolytic dissociation constant of 56 

molar depression of freezing-point 6 

temperature coefficient of conductivity, dissociation 142 

Weak electrolytes 26 

Wijs, ionisation of water 37 

Wolf 157 

avidity of weak acids 99 

dissociation of concentrated solutions 132 

mixed solvents 157 

"Zeitschrift fur physikalische Chemie, ,T Ostwald 2